<<

Path Integrals: From to Photonics Charles W. Robson,1 Yaraslau Tamashevich,1 Tapio T. Rantala,2 and Marco Ornigotti1, a) 1)Photonics Laboratory, Tampere University, P.O. Box 692, FI-33014, Tampere, Finland 2)Department of , Tampere University, P.O. Box 692, FI-33014, Tampere, Finland (Dated: 25 June 2021) The path integral formulation of , i.e., the idea that the evolution of a quantum system is determined as a sum over all the possible trajectories that would take the system from the initial to its final state of its dynamical evolution, is perhaps the most elegant and universal framework developed in , second only to the of . In this tutorial, we retrace the steps that led to the creation of such a remarkable framework, discuss its foundations, and present some of the classical examples of problems that can be solved using the path integral formalism, as a way to introduce the readers to the topic, and help them get familiar with the formalism. Then, we focus our attention on the use of path integrals in and photonics, and discuss in detail how they have been used in the past to approach several problems, ranging from the propagation of in inhomogeneous media, to parametric amplification, and quantum in arbitrary media. To complement this, we also briefly present the Path Integral Monte Carlo (PIMC) method, as a valuable computational resource for condensed physics, and discuss its potential applications and advantages if used in photonics.

I. INTRODUCTION proposing the Lagrangian as a more natural basis for a theory of quantum mechanics, as opposed to a Hamiltonian-based The path integral formulation of quantum mechanics, de- method which he argued was less fundamental due to its non- 45 veloped in the mid-twentieth century, is not only a remark- relativistic form . This led him to propose the key idea that able synthesis of several of the core ideas of theoretical a quantum mechanical transition amplitude for a particle is physics, but it is also a powerful computational technique for given by a factor controlled by the action along that 46 the analysis of a huge variety of physical systems in very particle’s path . This also led him to state, that the classical different contexts, such as quantum mechanics1–6, quantum path of a quantum system can be interpreted as resulting from field theory7–10, gauge field theory11,12, black hole physics6, the constructive interference of all such paths. In other terms, quantum gravity13,14, string theory15, topology6,16condensed the action S of a given system counts, de-facto, the number of matter physics17–23, statistical mechanics2,6,24,25, polymer of the path in units of Planck’s constant h, and, there- 47 physics5,6, financial markets5, optical communications26, fore, for each path, the phase of the related is given atomic physics27,28, spectroscopy29, light propagation in by exp(i2π S/h) = exp(iS/h¯). This allows the evaluation of turbid media30 and classical31–36 and quantum37–39 optics, the interference pattern of the particle dynamics, and essen- amongst others. tially represents what is nowadays commonly understood as 2,5 The underlying concepts of the path integral approach are path integral . sometimes considered difficult to grasp. Indeed, from a philo- The pivotal step in the development of the theory occurred sophical standpoint, its underpinnings are extraordinary – in in the 1940s, when Feynman formulated his version of quan- describing a system, all possible paths between its initial and tum mechanics, a ‘third way’ following the earlier, well- 4 final state must be taken into account mathematically. This known Schrödinger and Heisenberg alternatives . Based on idea of summing over all paths has previously been charac- integration over all paths between initial and final physical terised ontologically as everything that can happen does hap- states, with each path contributing an action-dependent phase pen40. as Dirac had proposed, Feynman’s formulation culminated As befits the subject, the history of path integrals is rather in his important 1948 paper entitled Space-time approach to 1 circuitous and does not follow a straight line from Feyn- non-relativistic quantum mechanics . man’s seminal work on the subject, published in 19481, to From a conceptual point of view, one of the most interest- arXiv:2105.00948v2 [quant-ph] 24 Jun 2021 the present day. Already in the 1920s, the mathematician ing consequences of path integrals is that they can provide a Norbert Wiener had developed a method to treat Brownian deep understanding of the relation between quantum and clas- and diffusion using a technique of integrating over sical mechanics, as the limit h → 0 emerges naturally from the 3 paths, with formal similarities to Feynman’s eventual con- formalism as the classical limit of the theory . struction but in a purely classical context41–43. During the Although the concepts underlying the path integral formal- early part of the following decade, the general ideas of de ism may at first appear quite alien, they are deeply profound. Broglie and Schrödinger, that waves can be associated with Even if its utility were limited, the beauty of the idea would particle dynamics44, motivated Dirac to publish a paper in the still merit a wide audience. The fact is, however, that the path Physikalische Zeitschrift der Sowjetunion (Physical Journal of integral has immense value as a practical computational tech- the Soviet Union) setting the stage for future developments by nique in the physical sciences, and it can be applied to solve problems in many diverse areas, even beyond physics. Two examples showing the remarkable breadth of its applicability are its use in quantum , where the “sum over histories” a)Electronic mail: marco.ornigotti@tuni.fi is interpreted as a sum over all different spacetime configu- 2 rations interpolating between an initial and final state of the density , which in space basis means finite closed loop universe13, and in financial market modelling, where the for- paths. Thus, in terms of path integrals each of the particles malism has proven useful, since the time dependence of asset propagate from a position in space in imaginary time back prices can be represented by fluctuating paths5. to the same position, the time period being inversely propor- The universality of the technique has allowed scientists tional to the temperature. Then, with Metropolis Monte Carlo to tackle many problems, and gain tremendous physical in- it is possible to sample particle paths with correct weight in sight into them. In statistical physics, for example, path in- predefined temperature and collect data enough for conver- tegrals conveyed the basic framework for the first formula- gence of expectation values of relevant operators. tion of the renormalisation group transformation, and they are This approach based on imaginary time propagation is largely employed to study systems with random distribution called Path Integral Monte Carlo method (PIMC). David of impurities6. In particle physics, they allow one to under- Ceperley and coworkers have carried out seminal develop- stand and properly account for the presence of instantons48. ment work and a sizable number of PIMC simulations of In quantum field theory they provide the natural framework to various many-particle systems ranging from superfluid He17 quantise gauge fields11,12,49. In chemical, atomic, and nuclear and neutron matter59 to and in extreme physics, on the other hand, they have been applied to vari- conditions60, including both bosonic and fermonic particles. ous semiclassical schemes for theory6. Through In recent years, one of the authors of this tutorial (TTR) path integrals, topological and geometrical features of clas- has significantly contributed to taking PIMC simulations to sical and quantum fields can be readily investigated and be new though simple quantum particle systems, such as small used to create novel forms of perturbative and nonperturba- atoms61,62, molecules63–66, a chemical reaction67 and quan- tive analysis of fundamental processes of Nature15,50,51. In tum dots68–70, with the ultimate goal of providing a more addition to that, path integrals allow one to re-interpret es- accurate description of their electronic structure and related tablished results, such as, for example, the BCS theory of properties including many-body effects, and how they change superconductivity7,52, from a novel, more insightful, perspec- with temperature62,71,72. In this context, PIMC has proven to tive. be a very reliable and excellent method to calculate the elec- Explicit, analytical solutions to problems formulated in tric polarisabilities of atomic and molecular systems, leading terms of path integrals, however, are scarce, and only available therefore to an accurate estimation of the optical properties for very simple systems, such as a free particle, or the ubiq- of both individual small quantum systems and collections of uitous harmonic oscillator2. The complexity of the path inte- them, in the form of dilute gases66,73,74. gral formalism, in fact, increases very rapidly to overwhelm- A different approach, based on a real, rather than imagi- ing levels of difficulties for many simple problems. As an nary, time path integral (RTPI) has been recently proposed as example of that, the simplest quantum system, i.e., a single a way to describe the full of a quantum hydrogen , required nearly 40 years to be fully solved in system at zero-Kelvin, and to also characterise the evolution terms of path integrals53. of its eigenstates69,73,75–78. A combination of PIMC and RTPI On the other hand, it is amidst complex and computa- therefore gives the possibility to have a comprehensive tool tionally challenging problems that path integrals show their to study the properties of complex systems and their classical true potential, providing a simple, insightful and intuitive and quantum evolution. This feature in particular might prove perspective on the physical principles regulating such pro- to be very useful not only in chemistry and condensed matter cesses. To do that, numerical techniques, such as Monte Carlo physics, where this technique fluorished in the past decades, methods19,23,54–58, and the computational power of modern but also as a viable mean to understand and design the prop- supercomputers are crucial to their successful implementa- erties of materials of interest for photonics. tion. A fully integrable simulation platform, that allows control Most of the practical applications to numerical simulations of both electronic and photonic properties of matter exactly, of quantum particles with path integral approaches are sys- without the necessity to revert to approximations or effective tems with finite temperature in thermodynamical equilibrium theories, in fact, would constitute a tremendous resource to- with one of the statistical ensembles24. Finite temperature wards the optimisation of integrated photonic systems. equilibrium involves dynamics of constituent particles and ex- It is interesting to notice, that throughout the last 30 years, change of with environment or heat bath. These are the path integrals have been used to describe several problems in central factors in with phase tran- classical and , such as the propagation of light sitions, conductivities and other processes related to interac- in gradient index media31, the estimation of the channel ca- tions between constituent particles. Typically, one assumes pacity of classical and quantum fiber-based communication canonical ensembles, where both the number of particles and networks26, parametric amplification37,38, light-matter inter- the volume they occupy are kept constant at a given tempera- action beyond the rotating wave approximation27, decoher- ture T, but other ensembles can be chosen where needed. ence and in nonlinear spectroscopy29, and the effect For a many-particle system in finite temperature, there is of retardation in radiative damping28. Path integrals have also no , but the mixed state can be described with a been employed to link the nonparaxial propagation of light , and it turns out that it can be written in terms with different models for quantum gravity32. All these works of path integrals in imaginary time2,24. The expectation val- share the common thread of employing nonrelativistic path ues of are then evaluated by using the trace of the integrals to calculate the (i.e., the Green’s func- 3 tion) of the electromagnetic field in different contexts, and use linear in the formalism, and introduces Feynman this information to solve the problem at hand. A different ap- diagrams. The results from these two sections are then intu- proach, based on path integrals in quantum field theory and itively and qualitatively generalised for the case of a vector Feynman diagrams, has been recently introduced as a viable field in Sect. IX, as a reference point for Sect. X, where these way to handle classical79 and quantum39 optical phenomena results are applied to the particular case of an electromagnetic in arbitrary media. field propagating in a dispersive medium of arbitrary shape. However, the benefits of path integrals in photonics, namely The last section of Part II, namely Sect. XI then presents two their ability to calculate both the properties of matter and its explicit examples, of how path integrals can be used in quan- with light in an exact way, without the need of any tum optics. The first example presents how to describe the approximation both on the matter and light side, and the new onset of spontaneous parametric down conversion (SPDC) in physical insight that this could bring to photonics, remains to lossy media through path integrals, while the second example date uncharted territory. deals with the calculation of the rate of In this tutorial, we aim at introducing the concept and meth- of a quantum emitter surrounded by a dispersive medium. ods of path integrals to the reader unfamiliar with the field, In the spirit of the educational purpose of a tutorial, and and to provide researchers in optics and photonics with a ref- given the mathematical complexity of path integrals especially erence point for both analytical and numerical methods in- concerning the concepts introduced in Part 2, we also provide, volving path integrals, with the hope that this will provide a in Appendix A, a step-by-step guide on how to deal with path powerful and practical toolkit that could be used in the fu- integral calculations for the explicit case of the electromag- ture to tackle challenging problems in photonics. An accurate netic field in arbitrary media. We hope this would serve as a description of the diverse interactions between light and mat- good reference and guide to better understand the techniques ter, emerging from the interplay of fundamental particles and and methods presented below. fields, calls naturally for the use of quantum physics, and its Finally conclusions and future perspectives are then given degree of complexity grows very quickly. Path integrals are a in Sect. XII. natural way to study these interactions, and they actively take advantage of the complexity of the problem. Using them in photonics might then lead to novel methods to exploit com- II. FUNDAMENTALS OF PATH INTEGRALS plicated light-matter dynamics in photonic systems. The tutorial is split into two main parts: Part 1, comprising A. Amplitudes: Classical vs. Quantum Sections II through VI, covers the basics of the path integral approach in physics, and presents examples on how they can Quantum physics is an abstract theory, whose specific fea- be used to solve problems in classical and quantum optics, as tures beyond are typically only spectroscop- well as how to employ PIMC to determine optical properties ically . A good starting point to find the underly- of materials. In particular, Sect. II covers the fundamentals of ing differences between the two seemingly different worlds of the path integral approach in physics, treating core concepts classical and quantum physics, is represented by their differ- such as the principle of least action, classical and quantum ent interpretation of the concept of probability. This, in fact, , and a brief description of the mathematics of in- turns out to be a direct manifestation of the wave nature of tegration over infinite number of paths. Basic examples, com- quantum particles, and thus, the fundamental issue that we prising the dynamics of a free quantum particle, the quantum need to incorporate in the study of the dynamics of quantum , and from a double slit are cov- systems. ered in Sect. III. To conclude Part I, two examples of the use The necessity of a change in viewpoint concerning proba- of path integrals in classical and quantum optics are presented, bility, and the consequent definition, for quantum physics, of namely the propagation of light in an inhomogeneous medium a complex-valued , emerges very clearly and how this could be related to the physics of a harmonic os- within the context of the least action principle. Let us consider cillator with a time-dependent , in Sect. IV, and the the situation depicted in Fig. 1, where a classical [panel (a)] investigation of degenerate parametric down conversion pre- and a quantum [panel (b)] particle are evolving from an initial sented in Sect. V (based on Ref. 37), respectively. Finally, time ta to a final time tb. Sect. VI briefly discusses how PIMC can be used to predict For the classical, deterministic, system, although many dif- the optical properties of matter, and presents some perspec- ferent paths joining ta with tb are available, only the stationary tives on the use of this computational resource for photonics. paths with least action [red line in Fig. 1 (a)] gives a signif- Part 2, on the other hand, including Sections VII through icant contribution to its dynamics, and determine, ultimately, XI, deals with the basics of path integrals in quantum field its equations of motion. theory (QFT), and presents an application of such framework For a quantum system, on the other hand, its intrinsic wave to the case of the dynamics of the electromagnetic field in ar- nature (and, ultimately the ) prevent it for bitrary media. In particular, Sect. VII briefly introduces the following one single path, and the classical path [blue tube concept of path integral for quantum fields, and makes use in Fig. 1 (b)] must be interpreted as the one with maximum of the simple case of a scalar field as an example to calcu- constructive interference coming from all possible paths. late the relevant quantities and establish the formalism. After To understand this better, let us consider a classical parti- having done that, Sect. VIII discussed how to include non- cle first propagating from a point xa to a point xc, and subse- 4

FIG. 1. Pictorial representation of the evolution of a classical (left) and quantum (right) system from an initial state at time ta, to a final state at time tb over a certain potential landscape. (a) While many different paths link the initial and the final state of the classical system, once the initial conditions have been fixed, only the path with minimal action (red line in the figure) is the actual one undertaken by the system. (b) For a quantum system, on the other hand, the classical path (blue tube) minimising the action is interpreted as the path along which the interference of all the contributing paths linking ta and tb is maximum.

quently to a third point xb. If we denote with P(c,a) ≡ P(xa → The probability amplitude ψ(b,a) defined above constitutes xc) the probability for the particle to propagate from xa to xc, the basic quantity from which the dynamics of a quantum par- and, similarly, with P(b,c) ≡ P(xc → xb) the probability for ticle can be derived, and it is usually referred to in literature the particle to propagate from xc to xb, the (classical) condi- as the kernel (or propagator, or Green’s function) of the quan- tional probability for the particle to propagate from xa to xb by tum system at-hand. The conventional symbol for that in path going through xc reads integral language is K(b,a), and we then adopt this notation for the rest of this manuscript. P(b,a) = ∑P(b,c)P(c,a), (1) c The kernel defined above has a simple physical interpre- tation. In fact, it can be thought as the impulse response of where the summation over c takes into account all possible the system at-hand2,81. Moreover, if we know the probabil- alternatives for the intermediate state c. Note, that the defini- ity amplitude of the system at a given initial state ψ(a), we tion of conditional probability given above, i.e., P(b,a), dif- can immediately evolve it to a final state characterised by a fers slightly from the usual one, which reads P(a|b). This probability amplitude ψ(b) using the following relation notation, however, is fully equivalent with the traditional one, and will turn out to be of more practical use for the purpose of Z this work. ψ(b) = K(b,a)ψ(a)dxa. (4) The above definition can be readily generalised to the case a of continuous variables, i.e., to probability densities, by pro- Finally, the experimentally observable classical probability moting P(a) and P(b) to probability density functions, and to distributions are found as squares of the absolute values of the interpret a and b as two sets of coordinates {xa} and {xb} for probability amplitudes, P(a) = |ψ(a)|2 and P(b) = |ψ(b)|2 80 the particle to occupy at a given time ta and tb, respectively . at times ta (initial state) and tb (final state), respectively. With In this case, then, the summation over all possible alternatives this definition, the probability amplitudes appearing above can c in Eq. (1) becomes an integral over the set of coordinates be readily interpreted as the wave function of the quantum sys- {xa}, i.e., tem. Equation (3), in particular, hints at the interpretation of the wave function of a quantum system as the sum (or, better Z said, interference) of all possible paths linking the initial and P(b) = P(b,a)P(a)dxa, (2) a final states of the considered evolution. where the subscript a on the integral indicates that the inte- gration has to take into account all the possible values of the B. Lagrangian, Action and Path Integral integrating coordinate xa. Now, let us extend the concepts introduced above to the case of a quantum particle with wave nature. To do that, let us Contrary to the canonical formulation of quantum mechan- first rewrite Eq. (1) for the probability amplitude ψ associated ics, which bases its premises on the Hamiltonian function H to the quantum particle as of the system, and therefore on the concept of total energy82, ψ(b,a) ≡ K(b,a) = ∑ψ(b,c)ψ(c,a). (3) the path integral formalism starts from the Lagrangian func- c tion, generally defined as the difference between kinetic and 5 potential energy of the system83, i.e., L = T −V. For this rea- and Planck’s constant h as the wave length. Then, we can as- son, the path integral formalism is often referred to as the third sign a wave exp[i2πS(x(t),a)/h] to the path x(t) and follow formulation of quantum mechanics, with the first being the the phase of the waves to find the interference effects. In par- developed by Heisenberg, Born and Jordan ticular, we want to take into account the contributions of all in 192584–86, while the second one being the familiar Hamil- possible waves of the form exp[iS(b,a)/h¯] to the probability tonian formulation developed by Schrödinger in 192687. amplitude K(b,a). Although the Hamiltonian and Lagrangian formulation of The sum (or integral) over the contributions of all possible physical problems are essentially, from the point of view of paths is called the path integral, i.e., physical meaning, equivalent, the latter is more elegant, and, thanks also to its natural appearance in the path integral for- b   malism, has been adopted as the natural framework for more Z i 8,88 89 K(b,a) = Dx(t) exp S(b,a) , (9) complicated theories, such as QFT , particle physics , and a h¯ string theory51. For simplicity of notations, we proceed in using the one di- where the notation Dx(t) indicates integration over all paths mensional space with the coordinate x, but the generalisation from a = (x ,t ) to b = (x ,t ), which, following Ref. 2, can to three dimensions is trivial. Then, for a particle with a a b b be defined as m in motion on the path x(t) with velocityx ˙(t) in a potential V(x,t) the classical Lagrangian is b Z 1 Z dx1 Z dx2 Z dxN−1 1 Dx(t) = lim ··· , (10) L(x˙,x,t) = mx˙2 −V(x,t). (5) a ε→0 A(ε) A(ε) A(ε) A(ε) 2 On a way to both finding the classical equation of motion and where A(ε) is a suitable normalisation factor that ensures the concurrently incorporating the wave nature of dynamics of the limit to properly converge (an example of it is given in the next particle, on the path x(t) from ta to tb we define the action section, but the explicit for of A(ε) might change depending on the problem at-hand), and comes from the discretisation Z tb ε S(b,a) = dt L(x˙,x,t). (6) of the time interval in the action S(b,a) in N finite points, i.e., t a tb − ta = Nε, with t0 = ta, and tN = tb. This, on the other In Lagrangian mechanics the action is a parameter related to hand, implies a discretisation on the paths x(t), which now the path length, whose optimisation will lead to the equations are defined as xk = x(tk) = x(t0 + kε), with x0 = xa = x(ta) of motion. This procedure, i.e., optimising the path x(t) such and xN = xb = x(tb). This discretisation procedure allows us that δS = 0 is called “the principle of least action", and leads to consider each of the N − 1 integrals above as standard Rie- to the Euler–Lagrange equation83,90 mann/Lebesgue integrals in the variable dxk. Then, when all N −1 integrals have been computed, the limit ε → 0, together d ∂L ∂L with the correct definition of the normalisation factor A(ε), − = 0. (7) R b dt ∂x˙ ∂x ensures convergence of the path integral a Dx(t), and justi- fies its definition as integration over all possible paths. No- If we now substitute Eq. (5) into the equation above, we find tice, moreover, that the same line of reasoning will allow us, the following differential equation in Sect. VII to define the path integral for fields. ∂V(x) The explicit form of the path integral above is usually de- mx¨ = − , (8) ∂x fined by the form of the potential term V(x) appearing in the Lagrangian (5). For some potential functions there are ana- which we immediately recognise as Newton’s classical equa- lytical, closed-form, exact solutions to the Eq. (9), but one tion of motion. This constitutes the fundamental ingredient needs to prepare for numerical methods with possible approx- for defining the classical and quantum probabilities as in Eqs. imations in more general cases, such as multi-dimensional or (1) and (3). However, for the dynamics of a classical sys- many-body problems. In what follows, we present some of tem, the integral approach above is redundant, and the tra- the exact . ditional approach based on Newton’s equation of motion is favourable one in most cases. Similarly, in quantum mechan- The above result, obtained for a simple one dimensional ics Schrödinger equation is the best approach for simple prob- system, can be readily extended to three dimensional and lems. However, there are several sophisticated cases where many particle systems. While the former is straightforward to Eqs. (3) and (4) are more practical and easy to handle. For work out, in the latter case, the quantum statistics of these more complicated scenarios, therefore, we need to find and bosons needs to be taken into account explicitly, which K(b,a) makes the problem of finding the correct generalisation of the the kernel , and the right way to do that is to incorpo- 91 rate in the “classical" path-based approach above, the informa- path integral to the many body case less trivial . tion of the wave nature of quantum particles, i.e., to introduce In addition to that, the path integral approach also allows interference between paths. for an easy way to simulate the evolution of the density matrix Following the ideas of Dirac45 and Feynman1,4, we con- for finite temperature equilibrium systems. An example of this sider the action S as a (classical) measure of the path length, will be discussed below in Sect. V. 6

III. BASIC EXAMPLES FOR QUANTUM PARTICLES Similarly, the action can be then written as the sum of three terms, namely the classical action Scl(b,a), the action rela- For the evaluation of the wave function from the integral tive to the deviation y(t), i.e., Sy(b,a), and the mixed action equation (4) we need to calculate explicitly the kernel from Smix(b,a). Because of the fact that y(ta) = 0 = y(tb), however, the path integral, i.e., Eq. (9). In this section we therefore con- all the terms which contain linear terms in y result in a vanish- sider the simplest kernels, with examples following the book ing integral. Thus, only the second-order terms in y give rise of Feynman and Hibbs2. to a nonzero contribution to the total action, which can now In case the integrand is an exponential of a quadratic func- be written as Z t tion (Gaussian integral) the kernel can always be evaluated b 2 2 using recursively the basic Gaussian integral formula81 S(b,a) = Scl(b,a) + ay˙ + byy˙ + cy dt. (17) ta r Z 2 π 2 Notice how S (b,a) does not depend on the deviation y(t), dxe−ax +bx = eb /4a. (11) cl a and therefore the corresponding exponential can be treated as a constant, with respect to the path integration Dy(t). The Another useful result to keep in mind is the fact, that for given Kernel can then be written in the following form xa and xb endpoints, the contributions from other than the clas- Z 0  i Z tb  sical path interfere destructively and vanish. Thus, only the K(b,a) = Dy(t) exp (ay˙2 + byy˙ + cy2)dt classical action Scl(xb,xa) contributes, leading to the major 0 h¯ ta simplification for the kernel, i.e.,  i  × exp Scl(b,a) , (18)  i  h¯ K(xb,xa) = F(tb −ta)exp Scl(xb,xa) . (12) R 0 h¯ where the notation 0 Dy(t) is reminiscent of the fact that all the paths y(t) obey the boundary condition y(ta) = 0 = y(tb). As long as the action involves path variables only up to the The path integral above, then, can be written as a function of second order, therefore, the exact propagators in the form the time interval (t −t ) solely, i.e., above can be factored out from the path integral, leaving at b a Z 0  Z t  most to calculate a prefactor of the form F(tb −ta). i b 2 2 To illustrate how one arrives at the result above, let us con- F(tb −ta) = Dy(t) exp (ay˙ + byy˙ + cy )dt . 0 h¯ ta sider a general quadratic Lagrangian in x andx ˙ of the form (19) 2 2 This, ultimately, allows us to write the kernel in the following L(x,x˙,t) = ax˙ + bxx˙ + cx + dx˙+ ex + f , (13) simplified form where {a,b,c,d,e, f } are (arbitrary, but well-behaved) time-  i  K(b,a) = F(t −ta)exp S (x ,xa) , (20) dependent coefficients. The action corresponding to this La- b h¯ cl b grangian is then the integral of Eq. (13) with respect to time between two fixed end points ta and tb, as given by Eq. (6). which is equivalent to that of Eq. (12). Let us now assume, that xcl(t) is the classical path between the specified end points, i.e., the path for which δS = 0 holds. We can then represent x(t) in terms of deviations from the A. Path Integrals for a Free Quantum Particle classical path xcl(t) by introducing the function y(t) as The first example concerns the simplest quantum system, x(t) = xcl(t) + y(t). (14) i.e., a quantum particle of mass m, freely propagating with- out experiencing any interaction. Following the assumptions This substitution means, that instead of defining a points on made in the previous section, we discuss the case of a one the path by its distance from and arbitrary coordinate axis, dimensional free particle. The generalisation to an arbitrary we measure instead the deviation y(t) from the classical path. number of dimensions can be readily done, since the dynam- Moreover, at each time t ∈ [t ,t ], the variables x and y differ a b ics of a free quantum particle in D dimensions can be seen as only by the constant x (t), and therefore dx = dy for each cl i i the product of the independent evolution of D one dimensional point t . In general, then, it follows that Dx(t) = Dy(t). Notice, i particles46. that as a consequence of Eq. (14), y(t ) = 0 = y(t ), as at the a b The Lagrangian of a free particle of mass m is given by endpoints the path x(t) coincides with the classical path xcl(t). If we then use the change of variables defined in Eq. (14), mx˙2 L(x,x˙) = , (21) the Lagrangian (13) can be written as the sum of three terms 2 as follows and the equation of motion deriving from the Euler-Lagrange L(x,x˙,t) = Lcl + Ly + Lmix, (15) equations (7) is simply mx¨ = 0.The correspondent action S = R tb L(x,x,t)dt ta ˙ can be readily calculated explicitly by means of where Lcl (Ly) is just Eq. (13) with {x,x˙} → {xcl,x˙cl} part integration, and has the following form ({x,x˙} → {y,y˙}), and m(x − x )2 S = b a , (22) Lmix = (2ax˙cl + bxcl)y˙+ (bx˙cl + 2cxcl)y. (16) 2(tb −ta) 7

] ��� in the expression above) give rise to the following term i im im S = (x − x )2 + (x − x )2 = (x2 + x2) ��� h¯ 1 2h¯ε 2 1 1 0 2h¯ε 2 0 im  2  ������� ���� + x1 − (x2 + x0)x1 , (26) [ � h¯ε and that, in particular, the first term does not depend on the -��� integration variable x1. Integrating the above quantity with respect to x1 then gives Z     ��������� -��� m i im 2 2 dx1 exp S1 = exp (x + x ) -� -� � � � 2πih¯ε h¯ 2h¯ε 2 0 m Z  im  � [���� �����] × dx exp x2 − (x + x )x  2πih¯ε 1 h¯ε 1 2 0 1 r   FIG. 2. Real (blue, solid line) and imaginary (red, dashed line) parts m im 2 = exp (x2 − x0) , (27) of the kernel K(xb,xa) of a free quantum particle, as given by Eq. 2πih¯(2ε) 2h¯(2ε) (30), as a function of x = xb −xa, for a fixed time interval tb −ta = T. For this plot, we assumed m = 1 = h¯, which corresponds to measure where to pass from the second to the third line we have em- time in square meters, rather than seconds. Moreover, T = 1 has also ployed the change of variables X = ix1 and used Eq. (11). been used. Next, we take into account the terms in the action depend- 2 ing explicitly on x2, i.e., S2 = S1 +(x3 −x2) , and we integrate with respect to x2. We can do so, by simply taking the result of the integral of S1 given in Eq. (27) and multiply it by where xa,b = x(ta,b). To calculate the path integral for a free particle, we need to consider all the possible paths the particle r   m im 2 takes from the initial state (t ,x ) to the final state (t ,x ). To exp (x3 − x2) , (28) a a b b 2πih¯ε 2h¯ε do that, we simply divide the time interval T = tb −ta into N smaller intervals of length ε = ti+1 − ti (such that T = Nε), and integrate again, this time over x2. The result is similar to 2 2 calculate the action Si corresponding to the particle evolution that of Eq. (27), except that (x2 − x0) becomes (x3 − x0) , within each single interval, such that and 2ε → 3ε. It is now clear, that we can solve the set of N integrals in Eq. (25) by recursively applying terms of the N N m form (28) and then performing Gaussian integration with re- S = S = (x − x )2, (23) ∑ i ∑ i i−1 x N − i=1 i=1 2ε spect to the variable i. After 1 steps, we are left with the following result with x ≡ x and x ≡ x . We then substitute this result into 0 a N b r   Eq. (9) and evaluate the path integral over the set of N distinct m im 2 exp (xN − x0) . (29) trajectories, i.e, 2πih¯(Nε) 2h¯(Nε)

Z b 1 Z Z Z If we now notice, that Nε = T = tb −ta, it is easy to see that the Dx(t) → N dx1 dx2 ··· dxN−1, (24) limit operation in Eq. (25) can be readily performed, leading a A(ε) us to the final form for the path integral of a free particle, i.e., p where A(ε) = (2πih¯ε)/m is a factor included to ensure the r  2  m im(xb − xa) integral to converge. This factor, however, is not merely a K(xb,xa) = exp . (30) normalisation factor, since it is complex, and therefore it con- 2πih¯(tb −ta) 2h¯(tb −ta) tributes to the overall phase of the path integral. This is dis- The functional form of the real and imaginary parts of the ker- cussed in great detail in Ref. 2. Finally, we take the limit nel above for a constant time interval tb −ta = T are shown in N → ∞, to arrive at the following expression for the propaga- Fig. 2. It is worth to comment this result a bit. From it, in tor of a free particle fact, we see how the quantum dynamics of a free particle (but, ! more in general, of an arbitrary quantum system) couples with 1 Z Z i N K(x2,x1) = lim dx1 ··· dxN exp Si its classical dynamics described by the action (22). However, N→∞ N ∑ A(ε) h¯ i=1 this result also allows to shed light on the essential difference (25) between quantum and classical dynamics. While the classical The integrals appearing above are Gaussian in the variables principle of least action localised the propagation of a particle xi, and can be then readily calculated one after another. To see on a specific trajectory x(t) (i.e., the path of minimal action), how, let us first calculate explicitly the integral with respect the propagator K(x(t),xa) describing the propagation of a par- to x1. Once we have this result, we can perform the other ticle from the initial state xa along the classical trajectory x(t), integrations in cascade in the same manner. yields instead a complex wave function ψ(x,t) delocalising in First, notice that the relevant term in the action, depending time during propagation (obviously within the constraints of explicitly on x1 (i.e., those obtained by setting i = 1 and i = 2 the Heisenberg principle). 8

where the coordinate xc is a free parameter to optimise fol- lowing the principle of least action. For the case of a refracting from interface C, therefore, the "minimum " from A to B is found by optimising the coordi- nate xc. Following Eq. (12), the kernel for the refraction of a quantum particle from a planar interface C is given, up to an inessential constant F ≡ F(tb −ta), by

im  K(x ,xa) = F exp [x˙ (x − xc) + x˙a (xc − xa)] . (32) b 2h¯ b b

We leave as an exercise for the reader to figure out the ex- plicit expression of the normalisation constant F ≡ F(tb −ta). In the equations above, the two different constant x˙a tox ˙b of the photon follow from two different refractive in- dices. In optics, the optimisation of the optical path length is called Fermat’s principle. If we consider the problem from a perspective, in fact, it is easy to see how FIG. 3. Schematic representation of a ray of light being refracted the optimisation of xc leads directly to the celebrated Snell’s from a planar interface. A light source located at A emits a photon, 92 which gets detected at point B by a suitable detector. Between A and law of refraction , i.e., B, a planar interface C is placed, separating the two different media in which the photon propagates. In the language of path integrals, na sinθa = nb sinθb, (33) refraction can be interpreted as the necessary change in of the photon along the path where the action has the minimal value, where na,b = 1/x˙a,b are the refractive indices of the two media with respect to the contact point xc on the planar interface C. separated by interface C (that can be expressed as the inverse velocity of the particle in each side of C), and θa,b the angle the optical rays emerging from A and B, resepctively, make with it. The angle difference between the two sides of the B. Refraction of at an Interface interface C is reminiscent of the different wavelength that a photon with velocityx ˙a and one with velocityx ˙b experiences. With the above notations, consider now the space divided In this simple example, the classical and quantum ap- in two parts by a planar interface C, and let A and B be two proaches give identical explanations for observations. So, we points located at opposite sides of the interface, as it is shown see that the "quantum corrections", though not absent may in Fig. 3. Then, we assume constant but different potentials give not only small, but even vanishing contribution. at opposite sides of the interface. At both sides we expect the path of the photons to be those of a free particle. With this assumption, we can project the path connecting the points A C. Diffraction from a Double Slit and B onto a one dimensional subspace, for simplicity. Notice, however, that despite we will perform the calculations in this This is the well known “classic experiment" for demonstrat- one dimensional subspace, we still need to think in terms of ing the wave nature of light, or, if conducted with quantum three dimensional space when considering deflection of light particles like electrons, to expose the wave nature of particle rays from the interface C. dynamics93. The results of this section can be then thought Without loss of generality, we can assume that total energy as valid for both a photon, or a quantum particle. Experi- is conserved in both separate sides of the interface, and while mentally, to prove the wave nature of light one would need passing through it. However, since the two sides might have to perform this experiment with monochromatic light, while different values of the potentials (i.e., different refractive in- mono-energetic particles are needed to unravel the wave na- dices), the particle passing through the interface C needs to ture of particle dynamics. From a path integral perspective, change its velocity fromx ˙a tox ˙b, to account for this variation the interference pattern typical of such experiments naturally in potential energy. The change of velocity, moreover, occurs arises when the possible paths the particle can take to traverse at some position xc on the interface. the double slit are explicitly taken into account. As shown in Fig. 3, in case the straight line from A to B is Our first task is then to construct the kernel for this prob- not perpendicular to the interface, the observed path becomes lem. To this aim, let us consider the situation depicted in Fig. deflected at xc. 4, where a particle is emitted from a suitable source located at Now, we can write the action for the path as point A and propagates to a detector (or a collection thereof) placed at B, through a screen C with two slits C1 and C2 carved Scl(xb,xa) = Scl(xb,xc) + Scl(xc,xa) in it. Without loss of generality, we can assume the evolution m of the particle is that of a free particle, and that the only po- = [x˙ (x − x ) + x˙ (x − x )], (31) 2 b b c a c a tential it encounters is represented by the double-slit structure 9

FIG. 4. Left panel: pictorial representation of the double-slit experiment. A quantum particle is emitted from a source located in A and reaches a detector (or a distribution thereof) located at point B by passing through a screen C, on which two slits, C1,2 have been carved. Right panel: sketch of the various possible probability distributions that could be observed at the detector plane B. (a) Interference pattern P revealing the wave nature of the quantum particle. (b) P2 obtained by closing the slit C1. (c) Probability distribution P1 obtained by closing the slit C2. (d) Estimation of the total probability distribution P1 + P2, obtained by simply summing the distributions obtained in the cases (b) and (c), as would be valid for a classical particle, where no interference occurs.

(which, as a matter of fact, acts as a transfer function for the distribution P when interference pattern is observed (both slits particle). are open) [panel (a)], the probability distribution P2 observed To reach the detector at B, the particle takes a time τ to when the slit C1 has been closed [panel (b)], the probability P1 travel from A to the screen C (and, in particular, to one of the observed when the slit C2 has been closed [panel (c)], and fi- slits), pass through the screen, and then arrive at B after a time nally the probability distribution P1 +P2 obtained by summing T −τ, where T = tb −ta is the total time the particle takes to go the results of observations in (b) and (c). from A to B. However, since we don’t know exactly the time at which the particle arrives on the screen, we need to integrate As it can be clearly seen, the probabilities do not sum up, over all possible times. In other terms, since we don’t know as P 6= P1 + P2. If, on the other hand, we first sum the prob- the exact path the particle will take to reach C from A, we need ability amplitudes, i.e., ψ = ψ1 + ψ2, and then calculate the to integrate over all possible paths it might undertake. With probability distribution as P = |ψ|2 we get this in mind, and by remembering that a = (xa,ta), b = (xb,tb), and c1,2 = (xc1,2 ,tc1,2 ) represent the position and time at points A, B, and C1,2, respectively, the kernel for the propagation of a particle through a two-slit screen is given by Z 2 2 ∗ h P = |ψ| = |ψ1 + ψ2| = P1 + P2 + 2Re{ψ1ψ2 }, (35) K(b,a) = dτ K(c1,b; τ)K(a,c1; T − τ) i + K(c2,b; τ)K(a,c2; T − τ) , (34) where the first term accounts for the particle passing through where the third term is responsible for adding interference the slit C1, while the second one for it passing through the slit on top of the sum of probabilities P1 + P2 in Fig. 4 (d), thus C2. From the expression above, it is clear that interference leading to the correct result of Fig. 4 (a). The reader familiar must occur, since the probability to detect the particle at point with wave theory would immediately recognise that the prob- B is then proportional to |K(b,a)|2. ability distributions shown in Fig. 4 (a) and (d) occur as well In fact, we can arrive at the same conclusion by consider- for classical waves. Specifically, the distribution (a) occurs for ing the case above in terms of the probability amplitudes and the interference of correlated (i.e., first-order coherent) classi- observed probabilities in Sect. II A. In particular, let us have cal waves, while distribution (d) results from incoherent clas- a look at the probability distribution of the particle observed sical wave mixing. This is yet another indicator of the wave on the screen at B for different cases, namely the probability nature of quantum particles94. 10

] ��� we can then factor the propagator K(x2,x1) for the harmonic oscillators into two terms as follows ���  i  K(xb,xa) = F(T)exp Scl , (39) ������� ���� h [ � ¯ S = R tb dt L(x ,x ) where cl ta cl ˙cl , and the explicit expression -��� of the function F(T), which depends upon the time interval solely, is -��� Z  i Z T m  ��������� F(T) = Dy(t) exp dt (y˙2 − ω2y2) . (40) -� -� -� � � � � h¯ 0 2 � [���� �����] Because of our choice of boundary conditions for the devia- tions y(t) andy ˙(t), namely that they must both be zero at the endpoints, we can significantly simplify the calculation of the FIG. 5. Real (blue, solid line) and imaginary (red, dashed line) parts of the kernel K(x,0) of a harmonic oscillator, as defined by Eq. (52), path integral above if we allow the various paths y(t) to be assuming that the initial condition for the oscillator is xa = 0, and for represented as a as a given time interval T = 1. Analogously to the convention adopted ∞ nπt  in Fig. 2, m = 1 = h¯ have been assumed to plot this figure as well. y(t) = ∑ an sin . (41) n=1 T

The requirement that y(ta) = 0 = y(tb) andy ˙(ta) = 0 = y˙(tb), in fact, corresponds to say that both y(t) andy ˙(t) are T-periodic D. Path Integral for the Harmonic Oscillator functions, that can then be represented in a Fourier series. This representation of y gives possibility to specify a path As our next example, we consider a simple harmonic oscil- through the coefficients an instead of values of y at any par- lator, described by the following Lagrangian ticular time t. This can be seen as a linear transformation of coordinates, whose Jacobian J is a dimensionless constant in- mx˙2 mω2x2 L(x,x˙) = − , (36) dependent of ω, m or h¯. Moreover, we don’t really need to 2 2 evaluate J explicitly, since we can always recover the correct where m is the mass of the oscillator, and ω its characteristic normalisation factor at the end of our calculation, by requiring frequency. Using the Euler—Lagrange Eq. (7), that the equation of motion readsx ¨+ ω2x = 0, and the action S = r m t lim F(T) = , (42) R b L(x,x,t)dt 2 ta ˙ is explicitly given by ω→0 2πihT¯ i.e., that in the limit of → 0, where the Lagrangian of the m  2 2  ω S = (x − x ) cos(ωT) − 2xax ) , (37) 2Tsinc(ωT) a b b harmonic oscillator reduces to that of a free particle, we find the appropriate normalisation coefficient for a free particle. where T = tb −ta, and sinc(x) = sin(x)/x. For this reason, we omit J from the following calculations, To calculate the path integral for the harmonic oscillator, we and we restore the correct normalisation factor only at the very take a slightly different approach, than the one taken above, end of them. which allows us to perform calculations in an easier and more Before substituting Eq. (41) into Eq. (40), a couple more intuitive manner. Let us then assume that xcl(t) andx ˙cl(t) assumptions are needed, in order to easily compute the path represent the classical path of the oscillator, and its veloc- integral. First, we truncate the Fourier series to a finite number ity, respectively. We can then express any other possible path N, so that taken by the oscillator, as a deviation from the classical path, Z 1 Z Z Z Dy(t) → da da ··· daN, (43) i.e., x(t) = xcl(t) + y(t) andx ˙(t) = x˙cl(t) + y˙(t), and choose A(ε)N 1 2 the appropriate boundary conditions on y(t) andy ˙(t), i.e., and we can calculate it recursively as we did for the case of a y(ta) = 0 = y(tb) andy ˙(ta) = 0 = y˙(tb) as required by the prin- ciple of least action. free particle. We will then take the limit N → ∞ at the end of If we apply this change of variables to the Lagrangian the calculation. above, we get three different terms, namely Putting everything together, we obtain the following ex- pression for the term F(T), after performing the trigonometric 81 L(xcl,x˙cl) + L(y,y˙) + m(x˙cly˙+ ωxcly), (38) integrals , as Z Z ( where the first term has the form given by Eq. (36) and it rep- 1 im T F(T) = da1 ··· daN exp (44) resents the Lagrangian for the classical trajectory, the second A(ε)N 2h¯ 2 term is the Lagrangian (36) for the deviation, and the third ) N πn2  term amounts to a vanishing term (for the Euler—Lagrange − ω2 a2 . (45) 2 ∑ n equations) and can therefore be neglected . Using this result, n=1 T 11

Notice, once more, that the integrals above are all Gaussian IV. AN EXAMPLE FROM CLASSICAL OPTICS: PATH in the integration variables an. As we did for the case of the INTEGRAL DESCRIPTION OF LIGHT DYNAMICS IN AN free particle, therefore, we can perform them individually, and INHOMOGENEOUS MEDIUM then obtain the final result recursively. The result of a single integration over an is then given by We now take a look at how path integrals can be used to solve problems outside quantum mechanics, and apply this s Z   formalism to describe the propagation of the electromagnetic dan imT 2 2 exp fn(ω,T)a = , (46) A 4h¯ n T f (ω,T) field inside a weakly inhomogeneous medium, using the case n of gradient-index (GRIN) media as explicit reference. This problem has been solved by C. Gõmez-Reino and J. Liñares where in 198731. In their work, Gõmez-Reino and Liñares first rep- 2 2 resent the electromagnetic field in a GRIN medium as a su- n π 2 fn(ω,T) = − ω . (47) perposition of optical rays, and use this picture to calculate T 2 the propagator as a path integral over the rays’ trajectories.

Since there are no linear terms in an in any of the integrals They then provide an explicit expression for it, parametrised above, the final result of the path integration will be propor- in the so-called paraxial and field rays of an arbitrary (parax- 92 tional to simply the product of N independent terms fn(ω,T), ial) optical system . one for each value of n. This allows us to write Here, we take a different approach, with which we want to show how the free propagation of light in a medium can be N 1 seen as, essentially, the evolution of a massive quantum par- F(T) ∝ ∏ . (48) ticle in a harmonic oscillator potential with a suitably defined p f (ω,T) n=1 n frequency, which, in general, can be z-dependent. We will We now need to take the limit of the expression above for identify such massive particle with a photon propagating in- N → ∞. To do that, let us first rewrite the product above in the side the medium, and all the possible trajectories the particle following way, using Eq. (47) can take as the possible optical rays linking the initial (z = 0) and final (z = z) propagation plane in the medium. We will −1/2 −1/2 then calculate the diffraction kernel by means of path inte- N 1 N n2π2  N n2π2  2 grals, essentially following the results of Sect. III D and show ∏ p = ∏ 2 − ω = ∏ 2 n=1 fn(ω,T) n=1 T n=1 T how our calculations naturally suggest a representation of the −1/2 diffraction kernel in terms of Hermite—Gaussian functions. N  ω2T 2  × 1 − . (49) Without loss of generality, and for the sake of simplic- ∏ n2 2 n=1 π ity of exposition, we consider light propagating in a 1 + 1- dimensional GRIN medium, characterised by the following The first product does not depend on and, together with the ω profile Jacobian and the terms p2/εT deriving from the various inte- 2 2 2 2 grations, can be collected into an overall normalisation factor n (x,z) = n0[1 − g (z)x ]. (53) C. The second product, on the other hand, admits the follow- ing limit where n0 = n(0,z) is the background index and g(z) is a smooth-enough function that describes the evolution of the re- − / s fractive index along the z axis. N  w2T 2  1 2 1 lim 1 − = . (50) N→ ∏ 2 2 ∞ n=1 n π sinc(ωT) A. Diffraction Kernel as a Path Integral Putting everything together and evaluating C = pm/2πihT¯ from the free-particle-limit (42) we get the final expression of Let us first assume paraxial propagation of light in a the term F(T) as medium described by the refractive index given by Eq. (53). In general, if we know the field distribution at an initial plane r m z = 0 to be E(xa,ya), we can calculate the field distribution at F(T) = , (51) 95 2πihT¯ sinc(ωT) a plane z > 0 by means of the diffraction integral Z and, after substituting this result into Eq. (39), we obtain the E(xb,z) = dxa K(xb,xa,z)E(xa,0), (54) final form of the path integral for the harmonic oscillator to be where K(xb,xa,z) is the diffraction kernel, i.e., the Green’s r m  i  function of the paraxial equation K(xb,xa) = exp Scl , (52) 2πihT¯ sinc(ωT) h¯ ∂K  ∂ 2  2ikn = − + 2k2 n n(x,z) K (55) 0 ∂z ∂x2 0 where Scl is given by Eq. (37). The real and imaginary pary of the kernel K(b,a) for the harmonic oscillator are shown in with the boundary condition that K(x,xa,z) → δ(x − xa) for Fig. 5. z → 0. If we imagine the electromagnetic field propagating in 12 the medium described by n(x,z) as a bundle of optical rays, trajectories as x(z) = xcl(z) + y(z), with the boundary condi- then the diffraction kernel can be interpreted as a path integral tions that the deviation y(z) is zero at the endpoints z = 0 and over all the possible trajectories of the optical rays contained z = z, and we can then write the propagator, in analogy with in the field as Eq. (39) as Z x   b i  i  K(xb,xa,z) = Dx(z) exp S(xb,xa,z) (56) K(xb,xa,z) = F(z)exp Scl(xb,xa,z) , (60) xa λ λ where λ = λ/2π = 1/k, and the action functional S for an where optical ray propagating in the medium described by Eq. (53) is defined as s i ∂ 2 Z z F(z) = Scl(xa,xb,z). (61) S = L(x,x˙,z)dz. (57) 2πλ ∂xb∂xa 0 Following this line of reasoning, and representing the classical The Lagrangian L for an optical ray propagating in a medium action Scl(xa,xb,z) in terms of the so-called paraxial (H1(z)) with refractive index n(x,z) can be written as follows and field (H2(z)) rays of a general optical system, we can ob- p tain a similar result, in our (1 + 1)-dimensional model, than L(x,x˙,z) = n(x,z) 1 + x˙2. (58) the (2 + 1)-dimensional result obtained by Gõmez-Reino and J. Liñares31, namely96 To justify our starting assumption, i.e., that the electromag- s netic field in the medium can be seen as a collection of rays, kn we can use different arguments. One possibility would be to K(x ,x ,z) = 0 exp(ikn z) b a iH (z) 0 represent the field in its plane wave components and consider 2π 1 each plane wave as an optical ray95. Another, more inspiring, " ˙ 2 2 # H1(z)xb + H2(z)xa − 2xaxb possibility is to notice that the paraxial equation (55) is for- × exp ikn0 ,(62) 2H (z) mally equivalent to the Schrödinger equation for a quantum 1 particle of mass n0 in a potential n(x,z), where the propaga- where the dot stands for derivative with respect to z and tion direction z plays the role of time, and k plays the role of h¯. H1,2(z) are two independent solutions of the following equa- Thanks to this formal analogy, we can identify a single opti- tion of motion cal ray as a (massive) photon propagating in the medium, and then easily understand as the diffraction kernel can be seen 2 H¨1,2 + g (z)H1,2 = 0. (63) as a path integral over all the possible trajectories that such quantum particle can take when evolving from the initial state This result, however, is not particularly insightful, and the to the final one. solution presented above (or its two-dimensional counterpart In the form above, the Lagrangian is of little use, since it presented in Ref. 31) is quite hard to intuitively link to known provides no analytical solution for the trajectory of the op- results. For this reason, we present below a much clearer and tical ray and, by extension, doesn’t really allow for an easy intuitive approach, which we hope will help the reader in ap- handling of the path integral in Eq. (56). To circumvent this preciating the universality and versatility of path integrals be- problem we can however assume that the medium is weakly yond quantum mechanics. inhomogeneous, so that ∆n = n(x,z) − n0  n holds, and we only consider rays propagating in a small region around the z-axis, which corresponds to assumingx ˙  1 (this is equiva- B. Paraxial Propagation as a Harmonic Oscillator lent to assuming that the paraxial approximation holds). With these assumptions, we can Taylor expand the square root in Let us go back to the analogy to the Lagrangian (59) and Eq. (58) and the refractive index profile obtaining, to the lead- calculate the path integral deriving from it. The fact that the ing order inx ˙ and x frequency of the oscillator is now z-dependent doesn’t really  2 2  allow us to repeat one-to-one the calculations in Sect. III D. x˙ 2 x L(x,x˙,z) ' n0 1 + − g (z) . (59) In particular, after we introduce the deviations y(z), we cannot 2 2 represent F(z) in terms of a Fourier series anymore, since now the oscillator frequency depends on z as well. Instead of doing Notice, that the optical Lagrangian is quadratic in both x and that, then, we just follow the same line of reasoning that we x˙, and can be therefore interpreted as the Lagrangian of a used to calculate the path integral for a free particle in Sect. (shifted) harmonic oscillator with mass n and z-dependent 0 III A, namely we divide the propagation “interval" Z = z − resonance frequency Ω(z) = g(z). Thanks to this analogy, we b z into N smaller intervals of length ε = z − z , such that can calculate the kernel K(x ,x ,z) in the same manner we a i+i i b a Z = Nε, so that we can write the action correspondent to the did for the harmonic oscillator in the previous section, with Lagrangian (59) as the difference that now we need to account for the fact that the frequency of the oscillator varies with z. In particular, we n0 2 n0ε 2 2 Si(xi,x ) = (xi − x ) − Ω x , (64) can employ the same trick of writing the components of the i−1 2ε i−1 2 i i 13

2 2 where x0 ≡ xa, xN ≡ xb, and Ωi ≡ g (zi). This allows us to φ(z) = γ(z) − γ(0), and the quantities s(z) and γ(z) are deter- approximate the path integral in Eq. (56) using Eq. (24), and mined from the solution of the following differential equation evaluate it first over a finite set of N trajectories, and then to get back to the actual result by taking the limit N → . d2 ∞ ξ(z) + Ω2(z)ξ(z) = 0 (71) By doing this, moreover, we gain the advantage, that in each dz2 infinitesimal interval ε, the frequency Ωi of the oscillator is constant, and so we can use the results for the standard oscil- where ξ(z) = s(z)exp[iγ(z)], and Ω(z) ∈ R. First of all, notice lator within each interval of length ε. If we then introduce the the similarity between this result and the form of the propaga- quantities tor for a harmonic oscillator, as given by Eq. (52). In fact, the result above reduces to the propagator of a harmonic os- n0 cillator with constant frequency ω by means of the substitu- β = , (65a) p ελ tion s(z) = n0/ω and γ(z) = ω z. We can then interpret  1  the propagation of light in a medium described by the re- α = β 1 − Ω2ε2 , (65b) i 2 i fractive index (53) as basically being given by the propagator of a harmonic oscillator with a suitably chosen z-dependent we can then rewrite the path integral in Eq. (56) as N nested frequency (which depends on the longitudinal properties of Gaussian integrals, i.e., the refractive index). The propagator, moreover, is uniquely   determined by the amplitude and phase of the solution of a iβ 2 2 KN(xb,xa,z) = exp xa + xb harmonic-oscillator-like equation of motion for ξ(z), where 2 the z-dependent profile of the medium determines, again, the Z " N−1 # 2 characteristic frequency of the oscillator. × dx1 ···dxN−1 exp i ∑ αnxn Following Ref. 97 we can make this connection appear n=1 more evident, by rewriting the propagator above in terms of × exp(−iβx0x1)···exp(−iβxN−1xN), (66) the eigenstates of the refractive index potential given by Eq. and then the diffraction kernel can be calculated as (53). This can be done by first rewriting the sinφ(z) and cosφ(z) as complex exponentials and then using Mehler’s 98 K(xb,xa,z) = lim KN(xb,xa,z). (67) formula N→∞ exp−(X2 +Y 2 − 2XYZ)/(1 − Z2) The explicit expression for K can be calculated by recur- √ N 2 sively applying the following Gaussian integral result to Eq. 1 − Z ∞ n (66)  2 2 Z = exp − X +Y ∑ n Hn(X)Hn(Y) (72) Z n=0 2 n! I = dx expiαx2 − (a + b)x with Z = exp(−iφ(z)), X = x pn γ˙ /λ, and Y = x pn γ˙ /λ r b 0 b a 0 a iπ  a2 + b2   ab  to transform the resulting expression in terms of Hermite poly- = exp −i exp −i . (68) α 4α 2α nomials and finally obtain (we generalise z = 0 to a generic z = z for convenience) This procedure will lead to the quite compact result a ∞ r ∗ aN  2 2  K(xb,xa,za,zb) = ∑ ψn (xa,za)ψn(xb,zb), (73) KN(x ,xa,z) = exp i pNx + qNx − aNxax , (69) b 2π a b b n=0

where ψn(x,z) are the eigenfunctions of a harmonic oscillator where aN, pN, and qN are quantities that depend on β and αi, and their explicit expression can be found in Ref. 97. We can with z-dependent frequency Ω(z), and their explicit expres- then use the result above and take its limit for N → ∞ to get sion is given as the final form of the propagator, which is given explicitly as s r     follows 1 n0γ˙ 1 ψn(x,z) = exp −i n + γ(z) s √ 2nn! πλ 2 ˙ ˙    n0 γaγb in0 s˙b 2 s˙a 2     r ! K(xb,xa,z) = exp xb − xa in0 s˙ 2 n0γ˙ 2πiλ sinφ(z) 2λ sb sa × exp + iγ˙ x Hn x , (74) ( " 2λ s λ in0 2 2 × exp (γ˙ x + γ˙ax )cosφ(z) 2λ sinφ(z) b b a where s(z) and γ(z) are solutions of Eq. (71). For light propa- gating in , g(z) = 0 (and, therefore Ω(z) = 0) and the #) p above expression of the propagator reduces to the well-known − 2 γ˙bγ˙a xaxb (70) result of the resolution of the diffraction kernel in terms of Hermite-Gaussian eigenstates of the paraxial equation99. The where the proper limit for the quantities aN, pN, and qN has expression above is moreover fully equivalent to that found been taken as instructed in Ref. 97 (also see the same dis- by C. Gõmez-Reino and J. Liñares31, but it gives more insight cussion for the simple harmonic oscillator in Sect. III D), on how path integrals can be used to easily solve complicated 14 problems in optics, such as the propagation of light in a GRIN medium, whose analytical solution, even in the simplest cases, in not really available. By establishing the analogy with a Input Nonlinear medium Output time-dependent harmonic oscillator, on the other hand, path ω integrals give a rather elegant and remarkably simple result, p which can be intuitively hinted at using simple arguments, χ₂ such as the propagation of light in free space. ωs ωi

V. AN EXAMPLE FROM QUANTUM OPTICS: PATH INTEGRAL DESCRIPTION OF DEGENERATE PARAMETRIC AMPLIFIERS

Path integrals, in the form defined by Eq. (9), have also been used to approach various problems in nonlinear optics. In this case, as we will see throughout this section, the integral over all the possible trajectories of the quantum particle will be replaced, in the formulation originally proposed by Hillery and Zubairy in 198238, by an integral over all possible config- urations in the complex plane spanned by coherent states. As an explicit example, we consider here the case of para- metric amplification, whereby a signal incident on an optically FIG. 6. Pictorial representation of optical parametric amplification. 100,101 nonlinear material is amplified . A pump field imping- A nonlinear material with second-order susceptibility (χ2) is im- pinged upon by pump photons at and signal ing on a material characterised by a χ2 nonlinearity (see Fig. ωp 6) can combine with a signal field, leading to a depleted pump photons at ωs. Due to the excitation of the medium to a virtual en- and an amplified signal (and an idler field, in order to conserve ergy level by the pump (shown at the bottom of the figure), the signal photon stimulates the emission of photons at ω and, due to energy energy). s conservation, produces an idler field at ωi. For the case of a degen- In this section, the contents of which based on Ref. 38, we erate parametric amplifier, the signal and idler fields have the same will then apply the tools presented so far in this tutorial to frequency, i.e., ωs = ωi. the case of the degenerate parametric amplifier, thus providing an example of how path integrals can prove useful to solve problems in quantum optics. If we introduce the notation |α,ti ≡Uˆ (t,0)|αi, we can rewrite For the sake of simplicity, let us consider a single mode of the expression above in the following form, which will prove the radiation field (generalisations of this formalism to multi- useful in the remainder of this section mode fields is covered explicitly in Ref. 38), and we represent the field using coherent states, a natural choice here as the K(αb,tb;αa,ta) = hαb,tb|αa,tai. (76) Hamiltonians under consideration will be expressed in terms † To understand why the form above is useful to our means, let of creation and annihilation operators (a and a, respectively). us show how the quantity above naturally appears when cal- This assumption, in particular, will allow us to define the path culating expectation values of operators and, more generally, integral in terms of coherent states, i.e., as a path integral in correlation functions, in the so-called P-representation. We the field’s . assume that at time t = 0 the density matrix of the field can be written as Z A. Propagator as Path Integral Over Coherent States ρ = d2αP(α)|αihα|, (77)

We begin by defining the time-evolution as where P(α) is the Glauber—Sudarshan P-function104,105. In ˆ U(tb,ta), which evolves the system’s state at time ta to the this representation, the expectation value of any operator Oˆ(t) ˆ 82 state at time tb by |ψ(tb)i = U(tb,ta)|ψ(ta)i . Denoting the in the can then be written as time-dependent Hamiltonian of the system by H(t), the time-  R t 0 0 Z evolution operator is then Uˆ (t ,t ) = Tˆexp −i b H(t )dt , 2 b a ta hOˆ(t)i = Tr[ρOˆ(t)] = d αP(α)hα|Oˆ(t)|αi. (78) where Tˆ is the time-ordering operator102. If the electromagnetic field is represented in terms of coher- Using the expression above in combination with the complete- ent states, we can readily give a definition of the propagator ness relation of coherent states, i.e., using the time-evolution operator defined above as follows 1 Z d2α |α,tihα,t| = 1, (79) K(αb,tb;αa,ta) = hαb|Uˆ (tb,ta)|αai, (75) π where |αi represent the set of coherent states, defined as the where d2α = d Reα d Imα spans the complex-plane defined eigenstates of the annihilation operator, i.e.,a ˆ|αi = α|αi103. by coherent states103, any normal-ordered correlation func- 15

† † tion of the form haˆ (t1)aˆ (t2)···aˆ(tN−1)aˆ(tN)i can be then ex- as follows: pressed in terms of the propagator (76). For the simple case Z (Z t ∗ ∗ of haˆ(t)i, this can be readily shown as follows b hαα˙ − α α˙ K(αb,tb;αa,ta) = Dα(τ)exp dτ t 2 Z a haˆ(t)i = d2α P(α)hα|aˆ(t)|αi ) ∗ i Z − iH(α ,α;τ) , (85) = d2α P(α)hα|Uˆ −1(t,0)aˆUˆ (t,0)|αi Z with the integration measure taken to mean an integral over 1 2 2 2 = d α d β P(α)β |K(β,t;α,0)| , (80) all coherent states parameterised by τ, with α(t ) ≡ α and π a a α(tb) ≡ αb. where to pass from the second to the third line we have em- ployed the completeness relation for coherent states. For cor- relation functions containing more creation and annihilation B. Propagator for Quadratic Hamiltonians operators, more propagators, calculated at different times, will appear in the expression above38. The propagator is then We now apply the above results, focusing on the class clearly an important quantity, to evaluate such expectation val- of Hamiltonians which are at most quadratic in the creation ues. and annihilation operators. This class of Hamiltonians is We now show how to write the propagator as a path inte- of particular interest in quantum optics, since it describes gral. To start with, let us assume the Hamiltonian of our sys- second-order nonlinear phenomena, within the so-called un- tem is normal-ordered, i.e., H = H(aˆ†,aˆ,t), and as we did in depleted pump approximation101, and it can sometimes also Sect. III A assume to divide the evolution interval tb −ta into be used to describe harmonic generation in third-order non- 101,106 N slices of length ε = (tb −ta)/N. We can this choice linear systems . directly into Eq. (76) by inserting N times the completeness The most general quadratic Hamiltonian can be written as relation (79) and set t → t = t + ε for each entry. If we do j j−1 † †  2  so, we then obtain H(aˆ ,aˆ,t) = ω(t)aˆ aˆ + f (t)aˆ + g(t)aˆ + h.c. , (86)

 N Z with f (t) and g(t) arbitrary, but well-behaved, time-dependent 1 2 2 K(α ,t ;αa,ta) = d α ...d αNhα ,t |αN,tNi functions. The reader might notice a similarity between the b b π 1 b b Hamiltonian above and that of a forced harmonic oscillator2. × hαN,tN|αN−1,tN−1i...hα1,t1|αa,tai.(81) For this class of Hamiltonians, the integrals appearing in Eq. (85) are all Gaussian, and can be performed using the meth- The quantities hα j,t j|α j−1,t j−1i can be readily evaluated us- ods described in Sect. III. In particular, to calculate the path ing the time-evolution operator defined above, and noticing integral with the Hamiltonian defined above, one first needs that since t j −t j−1 = ε  1, we can Taylor expand the time- to discretise the paths over the coherent states α (in the same evolution operator to obtain manner we discretised the trajectories for the harmonic oscil- lator in Sect. III), then express the exponential appearing in Z t j † Uˆ (t j,t j−1) ' 1 − iε dτ H(aˆ ,aˆ,τ). (82) Eq. (85) in terms of Re{α} and Im{α} explicitly, and notice, t j−1 that the correspondent integrals are Gaussian in both the real and the imaginary parts of α. After having calculated a single With this result at hand, the individual terms hα j,t j|α j−1,t j−1i term, one could then calculate the remaining integrals itera- can be brought, after a simple algebraic manipulation38, in the tively, as we have done for the examples in Sect. III. After a following form lengthy but straightforward calculation, which is partially cov- ered in Appendix A of Ref. 38, the propagator for a quadratic  2 2 ∗  hα j,t j|α j−1,t j−1i ≈ exp (−1/2) |α j| + |α j−1| + α j α j−1 Hamiltonian assumes the following form × exp−iεH(α∗,α ,t ), (83) j j−1 j−1 K(α,t) = F(α,t)exp[−iΣ(α,t)], (87) where we have defined where α = {αa,αb}, t = {ta,tb} and the functions F and Σ are hβ|H(a†,a,t)|αi defined as H(β ∗,α,t) = . (84) h | i h 1 β α F(α,t) = exp − (|α |2 + |α |2) +Y(t )α∗α 2 b a b b a Now that we have an expression of the propagator in terms i + X(t )(α∗)2 + Z(t )α∗ , (88) of N discrete “trajectories" (i.e., N different coherent states b b b b |α ji), we can take the limit N → ∞ and arrive at the definition Z t of the path integral in representation, following b n h 2 2 2 the same line of reasoning used in Sect. III. The details of this Σ(α,t) = dτ f (τ) 2X(τ) + Z (τ) + αaY (τ) ta calculation are reported in Ref. 38, and we refer the interested i h io reader therein. The final result of this calculation is then given + 2αaY(τ)Z(τ) + g(τ) Z(τ) + αa Y(τ) , (89) 16 where the auxiliary function X(t) is constrained by degenerate parametric amplifier:

 2 2  dX p |αb| + |αa| = −2iω(t)X − 4i f (t)X2 − i f ∗(t), (90) K(αb,tb;αa,ta) = sech[2κ(tb −ta)]exp − dt 2 ( ∗ −iω(tb−ta) with initial condition X(ta) = 0. The functions Y(t) and Z(t) × exp αb αae sech[2κ(tb −ta)] are instead defined in terms of X(t) as  α∗ 2 b −2iωtb  Z t  − i √ e tanh[2κ(tb −ta)] Y(t) = exp −i dτ [ω(τ) + 4 f (τ)X(τ)] , (91) 2 t ) a  2 i αa 2iωta − i √ e tanh[2κ(tb −ta)] . (95) and 2

Z t Z(t) = −i dτ [g∗(τ) + 2g∗(τ)X(τ)] ta VI. PREDICTING OPTICAL PROPERTIES OF MATTER  Z t  USING PATH INTEGRALS × exp −i dτ0 ω(τ0) + 4 f (τ0)X(τ0) . (92) τ The examples discussed above only deal with a single par- ticle, either freely propagating, or interacting with a suitable potential. In all these cases, the complexity of the problem is C. Propagator for Degenerate Parametric Amplification low enough, to allow analytical solutions to the path integrals. As soon as many-body effects are taken into account, as it is A degenerate parametric amplifier is characterised by the the case, for example, for and , handling the following quadratic Hamiltonian101 path integral analytically becomes impossible, and numerical techniques for its evaluation are necessary. A particularly suc- H(t) = ωa†a + κ e2iωt a2 + e−2iωt a†2, (93) cessful computational method for this task has been Path Inte- gral Monte Carlo (PIMC). The interested reader can download where ω is the angular frequency of the mode and κ is a cou- David Ceperley’s group PIMC++ open-source code for Path pling constant. The above Hamiltonian can clearly be seen to Integral Monte Carlo simulations, available, together with its fall into the class of Hamiltonians given in Eq. (86) if we iden- relative documentation, at https://pimc.soft112.com. tify ω(t) = ω, f (t) = κe2iωt , and g(t) = 0. Although para- In this section, we then present some of the results obtained metric amplification is formally a 3-wave process involving, using PIMC to predict the exact linear and nonlinear opti- cal properties of simple systems as a function of frequency. as depicted in Fig. 6, a pump, a signal, and an idler field, it 107 is common practice in nonlinear optics experiments to work These results, recently demonstrated by Tiihonen et al. , within the so-called undepleted pump approximation101,106, constitute one of the first attempts to use PIMC to investi- which treats the pump mode as a classical (bright) field, whose gate the optical properties of matter. There, however, only very simple systems were studied, such as the hydrogen (H) number of photons does not change significantly (i.e., the + + and hydrogen-like (Li and Be2 ) atoms, the atoms pump field remains undepleted) during the nonlinear interac- + + He and He , the hydrogen H2 and H2 , hydrogen- tion. Within this approximation, then, the Hamiltonian be- + + comes effectively quadratic in the signal and idler modes, and helium (HeH ) and hydrogen-deuterium (HD ) molecules, can be described by Eq. (93) above. The explicit form of the and the atom. These systems have been stud- propagator given below is then valid within this approxima- ied employing different methods, including finite-field sim- ulations for static polarisabilities65, polarisability estimators tion. A fully quantised analysis of parametric amplification 74 beyond the undepleted pump approximation in terms of path for simulation without the external field , static field-gradient polarisabilities66, and finally, dynamic polarisabilities and van integrals has been presented by Hillery and Zubairy in Ref. 73 37. der Waals coefficients . The last one, in particular, is of great interest, as the macroscopic electric susceptibility is con- In this case, with the form of ω(t), f (t), and g(t) given structed starting from the dynamic frequency dependent polar- above, the functions X(t), Y(t), and Z(t) can be written in an isabilities. explicit form (in particular, Eq. (90) cabn be solved analyti- The results presented in the aforementioned references, and cally), giving Z(t) = 0 and38 summarised in this section represent an exceptional bench- mark to gauge the capabilities of PIMC in providing accurate 1 −2iωt X(t) = e tanh[2κ(t −ta)], (94a) and multidimensional information about the optical properties 2i of matter, and hint at the potential impact PIMC could have in −iω(t−ti) Y(t) = e sech[2κ(t −ta)]. (94b) photonics, as a new modelling platform to calculate exactly the optical response of exotic materials, such as 2D materials, These time-dependent quantities can then be directly substi- the understanding of which is still in its infancy108. A great tuted into Eq. (87)), finally yielding the propagator for the advantage PIMC would give, for example, is the possibility to 17 investigate the nonlinear properties of 2D materials far from with real-valued, exponentially decaying terms, rather than equilibrium, a physical regime that is currently poorly under- complex-valued, spuriously oscillating ones. In fact, the ev- stood. This could be of particular importance for the nonlinear erywhere positive exponential function can be considered as a optical properties of 2D materials, since for such materials, probability distribution for sampling the imaginary time paths non-equilibrium dynamics can be easily reached at relatively with Metropolis Monte Carlo algorithm in canonical (NVT) small optical powers. ensemble. PIMC makes use of several different strategies to com- pute the quantities defined above. The zero-field polarisabil- A. Polarisability in PIMC ity estimators, for example, are Hellmann—Feynman type operators107, whose expectation values give the polarisabili- The optical response of matter95, frequently described in ties of various types and order, without the need of including terms of the matter polarisation vector P, is completely deter- an external electric field in the simulation. mined by dynamics solely within the limits of the the Computation of dynamic multipolar polarisabilities, on the Born-Oppenheimer (BO) approximation, which specifies that other hand, is more sophisticated and it requires to first de- the nuclear dynamics occurs on a much longer timescale and termine, through PIMC, the multipole-multipole correlation can be therefore neglected, it is implicitly assumed91. In gen- function in the so-called imaginary time representation, then eral, however, when thermal effects are explicitly taken into analytically continue it to obtain the spectral function in the account, this approximation might not be valid anymore, as real time domain (a task that is highly non-trivial, due to the the role of the nuclei becomes more prominent as the tem- fact that it is an ill-posed numerical problem), and finally both perature of the system is increased. To properly take into the real and imaginary parts of the polarisability. account these effects, then, an approach able to go beyond BO is needed to get the correct results. PIMC, then, is the most viable, if not the only possible, approach to systemati- B. From Polarisability to Susceptibility cally study the optical properties of materials in this particular regime (which, for example, includes the calculation of such Assuming a very low density gas phase of these small atoms properties at room temperature). or molecules, one can evaluate the corresponding macroscopic While optics makes extensive use of the (linear and nonlin- susceptibility from the atomic and molecular polarisabilities. ear) susceptibility tensor to describe the optical properties of a At higher densities, interactions and chemical reactions of material101, from a PIMC perspective, it is better to work with these moieties will change the polarisabilities, composition polarisabilities, as they can be defined quite easily in terms or both, which consequently leads to the density dependent of path integrals. For example, the i j-component of the lin- susceptibilities. ear (i.e., ) polarisability tensor, namely αi j, can be sim- With increasing density and, in particular, in case of liq- ply calculated as the (functional) derivative of the expectation uids and solids, where several electrons interact strongly, the value (i.e., first order correlator) of the dipole moment µ ori- Sign Problem (FSP) emerges. This is the notorious ented, say, along the i-direction, with respect to an external challenge for Monte Carlo evaluation of essential expectation electric field oriented along the j direction107, i.e., values of identical fermions, and it discloses partly a still open problem. The problem emerges from the evaluation of the dif- ∂ ference of two large contributions with opposite signs, result- αi j = hµii, (96) ∂E j ing from the sign of the density matrix (or wave function). There are practical solutions to FSP in PIMC simulations for where the expectation value of an operator Oˆ is defined with many cases20–23,107, but not any general robust approaches, respect to the density matrix operator of the system in the so- yet. Most of the solutions are based on finding approximate called imaginary time representation107, i.e., or iterative nodal surfaces of the density matrix. The real-time path integral (RTPI) approach74–77 may of-   S(T) hOˆi = N Tr Oˆ exp − , (97) fer a remedy in the future, as it works directly with the wave h¯ function with explicit sign, even though only at zero Kelvin. However, as for most of the cases of interests, electrons at where N is a suitable normalisation constant, S(T) = room temperature can be well-described with the zero Kelvin  Hˆ0 − Qˆ · E /kBT is the temperature-dependent action of the temperature model, this is not an issue. Moreover, including system (described by the Hamiltonian Hˆ0) interacting with temperature in RTPI should be, in principle, possible. the electromagnetic field E, and Qˆ is the electric moment Overall, the FSP represents a challenging problem, and sits operator107. Notice that to represent the expectation value at the forefront of current research in PIMC methods. The in- above as a path integral, it is computationally more conve- terested reader can find in Refs. 20–23, and references therein, nient to transform the complex phase factor exp(iS/h¯), nat- a good starting point to delve deeper in this fascinating, yet urally arising from path integrals, into an exponentially de- challenging, problem. In the examples that we are going to caying term exp(−S/h¯) by means of a Wick rotation109 ( present below, however, we have purposely selected simple i.e., a change in reference frame from real to imaginary time, systems, where there are no more than two electrons in each namely t → iτ), since computationally it is much easier to deal moiety, where at low temperatures we can assume the system 18

FIG. 7. Real (top) and imaginary (bottom) part of the dynamic polarisability as a function of frequency for Hydrogen (H, left) and Helium (He, right) atoms at T = 2000 K, within the BO approximation. In these plots, a blue line represents the dipole polarisability α1(ω), while a green line represents the quadrupole polarisability α2(ω). The black dots represent the reference simulations of polarisabilities calculated at T = 0 K. Dotted lines correspond to big time steps, i.e., ∆τ = 0.1 for Hydrogen, and ∆τ = 0.025 for Helium, while solid lines correspond to small time steps, namely ∆τ = 0.05 for Hydrogen, and ∆τ = 0.0125 for Helium. In PIMC, time steps are defined, in imaginary time, as ∆τ = 1/(MkBT), where kB is the Boltzmann’s constant, T is the temperature of the system, and M is the so-called Trotter number, which controls the accuracy of the discretisation of the propagator exp(−∆τ Hˆ ). In these plots, the frequency is measured in (1 Ha = 2 Ry = 27.2 eV). To get a better grasp of the span of the frequency axis of the plots above, a frequency of ω = 0.1 Ha corresponds to a wavelength λ = 456 nm, while a frequency of ω = 1 Ha corresponds to a wavelength λ = 45 nm. This figure has been taken from Ref. 107 with permission of the author.

to be in its singlet , thus enabling us to directly in optics95,101. In both examples below, the polarisability is label the light fermions with opposite spins and avoiding FSP calculated from the Kubo-like formula for the time-dependent entirely. optical susceptibility91

i χ(t) = Θ(t)h[Pˆ(t),Qˆ(0)]i = −Gret (t), (98) C. Some Examples h¯

In this section we present some results concerning simple where the square brackets denote a , the angled systems, such as one- and two-electron atoms and molecules, brackets denote instead thermal average [see Eq. (97)], Θ(t) obtained using PIMC. In particular, we focus our attention is the Heaviside step function81, Qˆ is a multipole operator on the electric polarisability, to show what are the capabili- (which represents dipole or quadrupole interactions, for the ties of this computational method, and how it could be used cases presented below), Gret (t) is the retarder Green’s func- to provide a new interesting platform for photonics, to cal- tion of the perturbation Pˆ and the latter represents the pertur- culate the optical properties of interesting materials, beyond bation generating the optical susceptibility. The link between the standard approaches and approximations frequently used susceptibility and polarisability is better understood in Fourier 19

FIG. 8. Real (top) and imaginary (bottom) part of the dynamic polarisability as a function of frequency for molecular Hydrogen H2 for various temperatures. The left panel presents the results for the dipole polarisability α1(ω), while the right panel contains the results for the quadrupole polarisability α2(ω). The black dots represent the reference simulations of polarisabilities calculated at T = 0 K. Dotted lines correspond to big time steps, while continuous lines correspond to smaller time steps. For these simulations, the time step has been adapted according to the chosen temperature, and it varies from ∆τ = 0.02 to ∆τ = 0.05. In these plots, the frequency is measured in Hartree (1 Ha = 2 Ry = 27.2 eV). To get a better grasp of the span of the frequency axis of the plots above, a frequency of ω = 0.1 Ha corresponds to a wavelength λ = 456 nm, while a frequency of ω = 1 Ha corresponds to a wavelength λ = 45 nm. This figure has been taken from Ref. 107 with permission of the author.

space, where the following relation is valid73 ical artefacts73. The results have been then compared with polarisability data available in literature, and obtained using a Z d Ω A(Ω) 110 χ(ω) = − = hα(ω)i, (99) T = 0 K approach, with methods other than PIMC . This π ω − Ω + iη reference is indicated by black dots in Fig. 7. For low fre- quencies, i.e., in the region 0 < ω < 0.4 Ha (corresponding to where the spectral function A(ω) is related to the sus- 95 0 < λ < 1.14 µm), PIMC simulations reproduce very well the ceptibility via the Kramers-Kronig relation , i.e., A(ω) = standard T = 0 K result. This, essentially, means, that for the Im{χ(ω)}. PIMC is then used to calculate the value of the , the eigenstates at T = 2000 K are essentially expectation value in Eq. (98) (or, equivalently, the retarded the same as those calculated at T = 0 K, and, as a consequence Green’s function), and from it both the real and imaginary of that, the low-frequency polarisability in Fig. 7 fits perfectly part of the polarisability are then calculated. Notice, more- the results obtained at zero temperature110. over, that the real part of the polarisability Re{α(ω)} corre- sponds to the actual optical response of the material, while its As the frequency is increased beyond ω = 0.5 Ha (corre- imaginary part Im{α(ω)} is, basically the spectral function sponding, for the case of Hydrogen, to its ionisation energy), A(ω). the zero temperature approach fails to faithfully describe the The first example we present is given in Fig. 7, which de- optical properties of the system. In this high frequency range, picts the real (top) and imaginary (bottom) part of the dipole i.e., ranging from the near to the deep UV, PIMC pro- (blue line) and quadrupole (green line) polarisabilities α1(ω) duces instead a more accurate and reliable result. This is an and α2(ω) for the case of atomic Hydrogen (left) and Helium important result. Usually, the models used in photonics to de- (right) at T = 2000 K. The simulations have been performed scribe the optical properties of various materials, are based using PIMC with two distinct characteristic time steps ∆τ, one on a zero temperature approach. Although this is perfectly large (solid green and blue lines in Fig. 7) and one small (dot- fine (and enough) for most materials, since the wavelength ted green and blue lines in Fig. 7), to rule out possible numer- at which non-equilibrium effects are starting to become non- 20 negligible is either in the far IR or the deep UV, this is not the extend this formalism to the case of quantum fields, where the case for 2D materials, which can be easily be driven out of building blocks for the path integral are not particles and all equilibrium in the wavelength range of interest for different the trajectories they can possibly take to evolve from the initial applications (typically ranging from THz to visible). In this to the final state, but rather all the possible configurations of a case, traditional, zero-temperature-based, methods would fail field, that serve the same purpose. to correctly describe the influence of out-of-equilibrium ef- A quantum field can be seen as a continuous collection of fects onto the optical properties of 2D materials, while PIMC, harmonic oscillators, one for each point in space where the on the other hand, are able to exactly account for these contri- field is defined. As such, then the results presented for the har- butions, as they naturally take these into account. monic oscillator in Sect. III D can serve as guidance to better The second example deals with a slightly more complicated understand this section, and, more generally, the properties of system, namely molecular hydrogen H2. Figure 8 depicts the a QFT. Some of the language and tools used here, however, are real (top) and imaginary (bottom) part of the dipole (left) and slightly different than those introduced above, and it is worth- quadrupole (right) polarisabilities α1(ω) and α2(ω), respec- while to introduce some key concepts and methods proper of tively, for several different values of the ambient temperature QFT before specialising our attention to the electromagnetic T. A comparison with reference simulations obtained at T = 0 field. To do that in the most beginner-friendly manner, we use K from Ref. 110 is also shown (black dots in Fig. 8). the case of a scalar field as a guidance to construct the path Notice, in particular, the deviation from the zero tempera- integral, and show how interactions can be introduced in such ture behaviour in the small frequency regime (insets in the top a theory, and what is the consequence of doing that. Then, part of Fig. 8). This discrepancy is due to the fact, that in the we will qualitatively and intuitively extend these results to the case of molecules, the nuclear contribution (mainly nuclear case of a vector field, and use those results to look at the lin- vibrations induced by finite temperature) cannot be neglected, ear and nonlinear dynamics of the electromagnetic field in an as it significantly changes the behaviour of the polarisability arbitrary medium. at low . This change amounts only to a shift of the This section follows closely the notation and line of rea- value of Re{α1(ω)} for the dipole polarisability (top let inset soning of Ref. 109. For a more formal discussion of path in- in Fig. 8), while it radically changes the form of Re{α2(ω)} at tegrals in field theory, we invite the interested reader to con- low frequencies, for the case of quadrupole polarisability (top sult Refs. 8,9,88,109. For the sake of simplicity, right inset in Fig. 8). Notice, moreover, as in this case tem- c = h¯ = me = 1 are implicitly assumed. If needed, these con- perature has also a significant effect in shaping the imaginary stants can be easily reinserted in the final results of calcula- part of the quadrupole polarisability at very low frequencies tions, by means of simple dimensional analysis. Moreover, (lower right inset in Fig. 8). summation over repeated indices (Einstein summation con- Although they show the optical properties of very simple vention) is also understood. systems, such as H, He and H2, the examples presented in this section offer good insight on the capabilities and poten- tial of using PIMC for calculating the optical properties of A. Path Integral for a Scalar even more complicated systems. The large frequency range PIMC can span (from THz to the deep UV) represents an To start with, let us then consider a free scalar field φ(x) enormous advantage to test the current models for suscepti- (where x is a shorthand for {r,t}), whose Lagrangian density bilities and polarisabilities of photonic materials, and could in is given by particular provide tremendous insight on the nonlinear prop- erties of exotic materials, such as epsilon-near-zero materials, 1 µν 1 2 2 L0(φ(x)) = η ∂µ φ∂ν φ(x) − m φ(x) , (100) for example, as it provides a single, ab-initio method to calcu- 2 2 late the optical response in an exact manner, without inserting where η µν = diag(−1,1,1,1) is the Minkowski flat space any approximation or without introducing different models for metric tensor109, and the subscript 0 indicated that the La- different frequency regimes. grangian is that of the free field, i.e., φ(x) is not interacting Besides that, the examples provided in this section also help with either itself, or its environment. the reader in getting a better grasp of the breadth of path in- Before going any further, it is instructive to discuss the tegrals, and how they do not only represent an elegant, alter- physical meaning of the parameter m appearing in the above native formulation of quantum mechanics, but also how their equation. In QFT, it is customary to give m the meaning of use in atomic and can be pivotal for the mass of the scalar field φ, as it could represent a scalar mas- accurate description of the properties of matter. sive boson such as the Higgs boson89, or even a scalar fermion in more complicated theories involving dark matter111. In op- tics, on the other hand, we frequently refer to the case m = 0, VII. PATH INTEGRAL FOR CLASSICAL AND QUANTUM as the electromagnetic field is composed by photons, which FIELDS are massless bosons89. However, when propagating inside a medium, the electromagnetic field acquires a mass, pro- In the previous sections, we have seen how it is possible to portional to the refractive index of the medium92. This can describe the dynamics of a quantum particle (or a collection be easily proven by recalling that the relation for thereof) in terms of path integrals. In this section, instead, we a monochromatic electromagnetic field propagating inside a 21 homogeneous medium characterised by a refractive index n i.e., φ ' φ(ti,x j,yn,zm) = φ(xk) ≡ φk (this implies that since is given by95 k2 = ω2n2. The dispersion relation for a mas- each field φ is a function of 4 variables, each taking N possible sive scalar field φ, on the other hand, can be readily cal- values, this discretisation of spacetime results in having a to- culated from its equation of motion (i.e., the Klein-Gordon tal of N4 fields). This then means, that the integration measure equation109), which can be readily derived from Eq. (100) can be written as using the Euler-Lagrange equations for fields, i.e., N N4 " # Dφ = lim dφ(ti,x j,yn,zm) ≡ lim dφk. (104) N→∞ ∏ N→∞ ∏ ∂L0 ∂L0 i, j,n,m=1 k=1 ∂µ  − , (101) ∂ ∂µ φ ∂φ In doing so, each field configuration φk becomes an indepen- R which leads to (∇2 − ∂ 2 − m2)φ = 0. If we now assume dent integration variable, and the integral Dφ reduces to a t 4 plane wave solutions for φ, the dispersion relation for a mas- usual Riemann/Lebesgue integral in N -dimensions. sive scalar field can be written as k2 = ω2 − m2. Notice, The above definition of path integral is valid both for a clas- now, that upon defining m2 = ω2(1 − n2), the dispersion re- sical, as well as a quantum field. For the latter, the field φ(x) lation for a massive scalar field φ matches that of an electro- needs to be equipped with a suitable set of commutation re- magnetic field inside a homogeneous medium. This allows lations, that define the quantum nature and structure of the us to identify the mass (squared) of the scalar field φ with algebra behind it. To do that, we can revert back to the anal- m2 = 1 − n2 = 1 − ε = χ, i.e., the medium’s susceptibility95. ogy with the harmonic oscillator, and use it to find a straight- In this sense, therefore, while a massless scalar field describes forward definition of such quantities. First, we discretise the a (scalar112) electromagnetic field in vacuum, a massive one field φ(x) and associate a harmonic oscillator to each point describes instead the (scalar) electromagnetic field in a homo- in space. Then, for each oscillator, we introduce the usual geneous medium. commutation relations between position and , i.e., We now proceed to the definition of path integral for the [qa(t), pb(t)] = iδab, where δab simply indicates that the oscil- scalar field φ. To do so, let us notice that the Lagrangian den- lators are mutually decoupled. If we then take the continuum sity in Eq. (100) has a similar structure to that of a harmonic limit (to revert back to the field φ(x)) the commutation rela- oscillator, as given in Eq. (36). The first term in Eq. (100), tions become those of the field φ itself with its canonically in fact, is quadratic in the derivative of the field, and can be conjugated momentum Π = ∂L /∂(∂t φ), namely associated with its “kinetic energy", while the second term,  0  0 quadratic in the field, is its “potential energy". Analogously φ(r,t),Π(r ,t) = iδ(r − r ), (105a) to what we have done in the previous sections, we can intro- φ(r,t),φ(r0,t) = Π(r,t),Π(r0,t) = 0. (105b) duce the action of the field as Equation (103), moreover, has implicitly been defined such Z 4 S0(φ(x),J) = d x [L0(φ(x)) + J(x)φ(x)], (102) that the time interval used for the time integration in the action spans the whole real axis, i.e., −∞ < t < ∞. Moreover, we 4 3 require that Z0(J) is normalised in such a way that Z0(0) = 1. where d x = dt d r and we have introduced the source term For a classical field, Z (J) represents its partition function J(x) (analogous to the external acting on a harmonic 0 2 Z0(J), from which, in analogy with , it oscillator ), which, for the sake of this tutorial, can be imag- is possible to derive all the relevant quantities of the classical ined to simply be a convenient mathematical tool, introduced field, such as energy, momentum, entropy, and Green’s func- to make calculations easier. The physical meaning of J(x), as tion. its name suggests, is to represent a source of the field. The For a quantum field, it represents the vacuum-to-vacuum x-dependence of the field φ(x) is from here henceforth, and correlation function, which allows to calculate transition prob- if not specified otherwise, implicitly assumed, i.e., we define abilities and correlation functions of field operators with re- φ ≡ φ(x), to make our notation less cumbersome. spect to the ground state (or vacuum), which is typically a sta- We can then define the path integral for a free scalar field in ble and well-defined state of the QFT at-hand. In other terms, the same manner as we did in Eq. (9), thus obtaining Z0(J) for a quantum field only describes a QFT at equilib- Z rium (i.e., at zero temperature). To deal with non-equilibrium iS0[φ,J] Z0(J) = Dφ e (103) systems, a further step in the definition of the path integral is needed, which is beyond the scope of this tutorial. The inter- where the functional measure Dφ(x) plays the same role as ested reader is then referred to Ref. 113 for further information the sum over all possible trajectories Dx(t) in Eq. (9), with the about non-equilibrium QFTs. conceptual difference, however, that in this case the sum (in- tegral) is extended to all possible configurations of the func- tion φ(x). Practically, this means that instead of slicing time B. The Propagator for a Free Scalar Field into small intervals (as we did in the previous sections for the case of quantum particles), we divide the volume of space- To get familiar with the formalism of QFT, in this section SN time V into N elementary cubes, i.e., V = k=1 Vk, with we calculate the propagator for the free scalar field, which 4 Vk = δ , and assume that in each Vk the field φ is constant, constitutes, to a certain extent, the QFT analogue of the free 22 quantum particle described in Sect. III A. To this aim, let us where first introduce the of the field as Z d4k eik(x−y) Z G(x − y) = , (111) ϕ(k) = d4xφ(x)e−ikx, (106a) (2π)4 k2 + m2 − iε Z d4k and a small imaginary part ε  1 has been introduced to reg- φ(x) = ϕ(k)eikx, (106b) (2π)4 ularise the ω-integral and allow its calculation as a contour integral in the complex plane81. The quantity G(x − y) is the where kx = k · r − ωt. If we substitute the expression of φ propagator (Green’s function) for the free scalar field, which above into Eq. (102) we arrive, after some straightforward cal- solves the correspondent equation of motion (derived from the culations, at the following expression for the action in Fourier Euler-Lagrange equations of the field), i.e., space 2 2 2 ∇ − ∂t + m G(x − y) = δ(x − y). (112) Z d4k h 1 S = − ϕ(k)k2 + m2ϕ(−k) 0 (2π)4 2 The Green’s function can also be interpreted as the vacuum- i to-vacuum two-point correlation function between the field at + J˜(k)ϕ(−k) + J˜(−k)ϕ(k) , (107) point x and the field at point y as follows109 where k2 = |k|2 − ω2, and J˜(k) is the Fourier transform of G(x − y) = ih0|Tˆφ(y)φ(x)|0i, (113) J(x). The trick to use here for calculating the path integral (103) with the action above is the same as the one we em- where Tˆ is the usual time-ordering operator. The two-point ployed in Sect. III A, namely we need to discretise the inte- correlation function is defined in terms of path integrals as gral appearing in the action S0, and consider a discrete set of follows R field configurations, such that Dφ can be written as a prod- Z uct of integrals. If we do so, we can then use the fact that h0|Tˆφ(x)φ(y)|0i = Dφ φ(x)φ(y) exp[iS0(φ,J)]. (114) the integrand above is Gaussian in ϕ(k) and repeatedly apply Gaussian integration to it to get the final result. From this definition, by noticing that, in analogy with the Alternatively, we can operate a change of path integra- derivative of the exponential of a function we can write5 tion, by performing the change of variables χ(k) = ϕ(k) − ˜ 2 2 δ J(k)/(k + m ), such that Dχ = Dφ, to isolate, in the action φ(y)eiS0(φ,J) = −i eiS0(φ,J), (115) above, a term that only depends on the field χ, from a term δJ(y) independent from it, i.e., where the symbol δ/δJ(y) stands for the functional deriva- 1 Z h i S = d4k − χ(k)(k2 + m2 − iε)χ(−k) tive, and it is a generalisation of the concept of derivative to 0 2 functionals5 defined by the operative relation 1 Z J˜(k)J˜(−k) + d4k . (108) δ f (x) 2 k2 + m2 = δ(x − y). (116) δ f (y) If we now substitute the expression above into Eq. (103), to- gether with Dχ = Dφ from the change of field variables, we With this in mind, then, we can formally introduce in Eq. obtain the following result (113) the substitution φ(x) → −iδ/δJ(x), thus obtaining an alternative definition for the propagator of a free field as  i Z d4k J˜(k)J˜(−k)Z n Z0(J) = exp Dχ exp Z 2 (2π)4 k2 + m2 h0|Tˆφ(y)φ(x)|0i = Dφ φ(y)φ(x)eiS0[φ,J] Z d4k  k2 + m2 o  2 2 − i 4 χ(−k) χ(k) . (109) 1 δ (2π) 2 = h0|0iJ i δJ(y)δJ(x) Notice, how the first term has been brought out of the path  δ 2  integral, as it is independent on χ. Notice moreover, that the = − Z (J) . (117) δJ(y)δJ(x) 0 path integral can now be evaluated, using a generalisation of J=0 the Gaussian integration formula (see Appendix A), and it es- sentially amounts to a normalisation factor (which, for con- venience, will be henceforth neglected). We then arrive at the VIII. NONLINEAR INTERACTIONS THROUGH PATH final result for the path integral, which can be written in a very INTEGRALS similar form to that of a harmonic oscillator, as The results above are suitable for describing the free prop-  i Z d4k J˜(k)J˜(−k)  agation of a field, such as, for example, the electromagnetic Z0(J) = exp 4 2 2 2 (2π) k + m − iε field in either free space (i.e., vacuum), or inside a homoge-  i Z  neous or inhomogeneous medium. However, neither the La- = exp d4xd4yJ(x)G(x − y)J(y) , (110) 2 grangian (100) nor the path integral (103) take into account 23 the possible self-interactions of the field, or any other form of If we now assume that the interaction is “small enough” interaction (such as the field-environment interaction, for ex- (with respect to a certain energy scale, typical of the prob- ample). In general, field interactions are very important, be- lem at-hand) that we can treat it perturbatively, i.e., if we set cause they allow us to gain valuable insight on how the field λ  1 we can expand the exponential in a power series in λ behaves in different environments. The self-interaction of a (truncated at the desired perturbation order) and write the path field with itself, in particular, plays a big role in photonics, integral for the interacting theory in the following manner where it is simply known as nonlinear optics. If we then want n V to use path integrals to describe quantum nonlinear optics, we ∞ 1 iλ Z 1 δ   Z(J) = d4x need a way to include the effect of interactions in our theory. ∑ V=0 V! n! i δJ(x) In QFT this is typically done by adding to the Lagrangian den- ∞ 1  i Z P sity a so-called interaction term, i.e., × ∑ d4yd4zJ(y)G(y − z)J(z) , (123) P=0 P! 2 L = L0 + Lint , (118) where we have also written Z0(J) in terms of its power series where Lint contains information about the nature of all the in- expansion, for a reason that will become clear in a moment. teractions taking place in the considered system. For the sake The expression above allow us to treat the path integral for of simplicity of exposition, and because they play an essential the interacting theory in a perturbative manner, up to an arbi- role in nonlinear optics, let us focus our attention on the case trary perturbation order in λ. To choose the appropriate one, of a self-interacting field, where the interaction part is poly- in fact, one needs only to truncate the summation in V at the nomial in the field φ and therefore the interaction Lagrangian desired order, so only terms proportional to λV (or lower pow- can be written as90 ers) will appear in the perturbative series expansion.

λ n Lint = φ , (119) n! A. Feynman Diagrams where λ is a suitable constant describing the strength of the interaction. Notice, that the first nontrivial interaction In general, calculating Eq. (123) analytically is not possi- is given by n = 3, as any quadratic term in the Lagrangian ble, and even writing down the terms contributing to Z(J) at density can be absorbed in the free Lagrangian as a “shift in a given perturbation order might be cumbersome. In fact, at the potential energy". order λV in Eq. (123), Z(J) contains nV functional derivatives Inserting Eq. (118) into Eq. (103) gives the formal expres- through the term (δ/δJ)n, that acts on the 2P different source sion of the path integral for the interacting theory, i.e., terms J deriving from the P−th order expansion of Z0(J). This, in general, amounts to (2P)!/(2P − nV)! different ways Z  Z  4 the nV functional derivatives can act on the 2P source terms. Z(J) = Dφ exp i d x [L + Lint + Jφ] . (120) 0 It is not difficult to see from this example, then, how the num- ber of required integrals to calculate becomes very high and In general, Z(J) cannot be evaluated analytically, because the complex very easily, with increasing V and P. integral at the exponent is intrinsically non-Gaussian, due to An elegant and quite insightful way to keep track of all the presence of the n > 2 polynomial terms provided by Lint . these quantities, and to focus on their physical meaning, rather However, we can use the following argument, to write Z(J) than their complex mathematical expressions is given by the in a form that can be treated with perturbation theory. First, so-called Feynman diagrams. An example of them for the case notice that the interaction Lagrangian Lint as defined in Eq. of n = 3 is reported in Fig. 9. (119) is a polynomial of order n in φ. Formally, then, if we To understand their meaning, a comprehensive set of rules employ Eq. (115), we can say that connecting the various parts of these diagrams with the cor- respondent integrals and mathematical quantities appearing 1 δ  Lint (φ) → Lint (121) in Eq. (123) can be defined, following for example Refs. i δJ 8,9,109, as follows: holds. • to represent a propagator −iG(x − y) we use a wiggly Now, thanks to Eq. (121), the interaction term Lint does not line joining the initial (x) point with the final point (y) depend anymore on the path integral variable φ, because we of the free evolution of the field φ; have formally substituted each occurrence of φ in it with the correspondent derivative with respect to J. This allows us to • to represent the interaction between fields, we use a so- bring this term outside the path integral, and therefore rewrite called vertex, i.e., a filled dot, which joins n lines, thus Eq. (120) as fulfilling the φ n nature of the interaction;  Z   4 1 δ • to each vertex, moreover, we associate the quantity Z(J) = exp i d xLint Z (J), (122) R 4 i δJ 0 iλ d x, which contains the λ defin- ing the interaction. In addition, we require that energy where Z0(J) is given by Eq. (103). and momentum conservation at each vertex holds; 24

the propagator. This corresponds, in terms of calculations, to the following

δ 2Z (J) < Tˆφ(x)φ(y) >= 0 δJ(x)δJ(y) J=0

J(x1) J(x2) ( ) δ 2 = δJ(x)δJ(y) J=0

(x) (y) = ≡ G(x − y). (125)

Notice, how the action functional derivatives δ/δJ(x), namely to remove a source term from Z0(J) and replace x1 3 FIG. 9. (a) Relevant for a scalar φ interacting (x2) with x (y), now appears more clear and intuitive. QFT for the simple case V = 1, P = 2. As it can be seen, the dia- gram contains one source term corresponding to the current J(z) in Eq. (126), and two propagators, G1 (black, wiggly line) and G2 (red, wiggly line), which is looping around the vertex. By exchanging the 2. First Order Perturbation Diagram for n = 3 roles of G1 and G2, the diagram doesn’t change, i.e., the value of Z1[J] in Eq. (126) doesn’t change. This diagram is known in QFT As a second example, let us consider a n = 3 interacting as a tadpole diagram. (b) Relevant Feynman diagram for the case of theory with first order perturbation (i.e., V = 1), for the simple V = 1 and P = 3. This diagram can be arranged in 3! = 6 differ- cases of P = 2, and P = 3 as reported in Fig. 9 (a), and (b), ent ways, by cycling through the currents J and the propagators 1,2,3 respectively. G1 (black, wiggly line), G2 (blue, wiggly line), and G3 (red, wiggly line), without changing the value of the correspondent partition func- The case P = 2 represents the simplest perturbative calcula- tion. This diagram will be of particular importance for the case of the tion that we can do on Z(J) for n = 3, and corresponds to the electromagnetic field, as its vector counterpart is the basic quantity well-known tadpole diagram in QFT, i.e., the case where two that describes second-order nonlinear processes in arbitrary media propagators interact at a single vertex V [Fig. 9 (a)]. Notice, (see Sect. X). that this diagram contains 2P − 3V = 1 external sources, and can be arranged in 2 different ways, by exchanging the role of the two propagators G1 and G2. Both of these configura- tions, however, are algebraically equivalent, and correspond to • finally, to represent the presence of sources J(x) we use the same term in the first order perturbative expansion Z1(J), a circle, placed at one end of a propagator, which is which for this simple case can be written, following the rules R 4 associated to the quantity i d xJ(x). above, as

With these simple rules, the terms at different perturbation  Z  3 i iλ 4 δ orders in Eq. (123) can be conveniently expressed in terms Z1(J) = − d x of diagrams. To illustrate how this works, let us consider two 4 3! δJ(x) examples: Z 2 × d4yd4zJ(y)G(y − z)J(z)

λ Z 1. Feynman Diagram for the Free Propagator = d4xd4zJ(z)G (z − x)G (x − x), (126) 2 1 2

To obtain the free theory from Eq. (123), we simply set where the last term G2(x − x) accounts for the loop appearing V = 0 (which corresponds to no interaction). Following the in Fig. 9 (a). The factor 1/2 in front of the integral com- rules stated above, the path integral Z0(J) can then be dia- pensates for the 2 equivalent ways diagram (a) can be written. grammatically written as follows The subscript 1 and 2 in the propagators have been added to make it easier to recognise the correspondent terms in Fig. 9, J(x1) J(x2) and have no physical meaning. Z0(J) = + higher orders. (124) For the case of three propagators meeting at a vertex (P = 3 and V = 1), the actual computation of Z1(J) would be rather If we then want to calculate the propagator, we just have to cumbersome, as it requires to perform three successive func- simply take the functional derivative of the diagram above, tional derivatives on an object that has six current terms, and i.e., remove the source terms J(x1) and J(x2) and replace them three propagators. If we however rely on the correspondent with simply the labels x and y for the initial and final points of Feynman diagram, depicted in Fig. 9 (b), we can readily write 25 the explicit expression of Z1(J) as where the shorthand J = 0 means that all the components of the current need to be set to zero after the calculation. We λ Z h Z (J) = d4xd4yd4zd4x J(x)G(x − x ) will make use of this result in the next section, for deriving 1 6 1 1 i the propagator of the effective free electromagnetic field in an G(x1 − y)J(y) + G(x − z)J(z) . (127) arbitrary medium. The diagram in Fig. 9 can be arranged in 3! = 6 different ways, by interchanging the roles of the three source terms Jk. All these different arrangements of the diagrams, however, are B. Path Integral for Interacting Vector Fields algebraically equivalent, and correspond to the same form of the first-order perturbation Z1(J) written above. The factor 1/6 there, moreover, accounts for this . Analogously, Eq. (123) can be qualitatively extended to the case of interacting vector fields by promoting source terms to vector fields (i.e., J → Jµ ) , and the propagators to tensor fields IX. GENERALISATION TO VECTOR FIELDS (i.e., G(x − y) → Gµν (x − y)), thus obtaining

The results obtained in the previous section for a scalar field V ∞ 1 iλ Z 1 δ n are valid, if properly generalised, also for vector fields, and, Z(J) = ∑ {α} d4x in particular, for the electromagnetic field. The correct way V=0 V! n! i δJα (x) to generalise those results to vector fields, however, is out of ∞ 1  i Z ←→ P the scope of this tutorial. The interested reader is referred to × d4yd4zJ(y) G (y − z)J(z) ,(132) ∑ P! 2 Refs. 2,5,8–10,88,109 for further and more formal details. In P=0 what follows, we just give an intuitive and qualitative gener- alisation. where λ{α} represents a rank-n tensor, whose indices are con- tracting the n indices appearing in the functional derivatives ←→ and J(y) G (y − z)J(z) = J (y)G (y − z)J (z). In general, A. Path Integral for Free Vector Fields µ µν ν moreover, λ{α} could also be space-dependent. This result will be used in the next section as a starting point to analyse For a vector field, in fact, Z0(J) can be qualitatively written nonlinear quantum effects for photons propagating in arbitrary in a similar manner to Eq. (110). To do that, we assume media. that the source terms are now vectorial in nature, i.e, J → Jµ , which makes the Green’s function become a two-index object, i.e., we need to formally make the substitution G(x − y) → Gµν (x−y). This allows us to immediately extend the result of Eq. (110) to the case of a vector field, by simply introducing X. PATH INTEGRAL FORMALISM FOR CLASSICAL AND IN ARBITRARY MEDIA the dyadic Green’s function   1 0 0 We now use the formalism introduced above for describing Gµν (x − y) = δµν + 2 ∂µ ⊗ ∂ν G0(r,r ;t,t ), (128) k the propagation and interaction properties of an electromag- 0 0 netic field in a dispersive medium of arbitrary geometry. The where G0(r,r ;t,t ) is the Green’s function of the wave equa- tion, i.e., content of this section follows the line of reasoning of Refs. 39,79, and it is reported here with a higher level of detail, in 2 2 2 0 0 0 0 ∇ − n ∂t G0(r,r ;t,t ) = −δ(r − r )δ(t −t ), (129) order to facilitate the reader in the process of adapting the for- malism developed in the previous sections to the case of the where n2 is a possibly space-dependent parameter describ- electromagnetic field. ing the properties of the vector field. For the electromag- Let us then start by making a series of assumptions to con- n2 netic field, for example, is the refractive index profile of struct a suitable Lagrangian density, which is the starting point the medium the field is propagating through. for constructing the path integral. First, we choose to work in We can then write the partition function for the non- the so-called Weyl gauge114 where the scalar potential is zero interacting vector field simply as (which corresponds to assuming that the medium we consider  i Z  contains no free carriers). This, practically, corresponds to the Z (J) = exp d4xd4yJ (x)G (x − y)J (y) . (130) 0 2 µ µν ν usual assumption that both the electric and magnetic fields are fully determined by the vector potential, i.e., E = −∂A/∂t Using Eqs. (113) and (117) we can also write Gµν (x − y) as a and B = ∇ × A. Then, we assume that light-matter interac- two-point correlation function as tion can be fully described within the framework of the dipole 115 approximation . δ 2 Gµν (x − y) = −i Z0(J) , (131) We can then describe the evolution of an electromagnetic δJµ (y)Jν (x) J=0 field in a dispersive medium with arbitrary shape in terms of 26 the Huttner-Barnett Lagrangian density116 a medium macroscopically described by an effective dielec- tric constant. We call the system electromagnetic field plus   ε0 2 1 2 g(x) ˙ 2 2 2 effective medium dressed electromagnetic field. LHB(r,t) = E − B + 2 P − ω0 P 2 2µ0 2ε0ω0 β(x) The full procedure on how to do that can be found in Refs.  2  Z ρ 2 ρω 39,79, and it is summarised in Appendix A, for completeness. + g(x) dω F˙ (ω) − F2(ω) 2 2  Z  − g(x) E · P + dω f (ω)P · F˙ (ω) . (133) A. Dressed Free

Let us first understand what is the physical meaning of all After performing the path integration with respect to P and F(ω) as illustrated in Appendix A, the partition function in the elements in the expression. LHB describes the interac- tion of three fields: the electromagnetic field (represented by Eq. (134) reduces to an effective partition function Ze f f (J) E ≡ E(r,t) and B ≡ B(r,t)), the matter (polarisation) field given by Z  Z  P ≡ P(r,t), and the reservoir field F(ω) ≡ F(r,t;ω), which 4 essentially models all the possible decay channels present in Ze f f (J) = D Aexp iSe f f (A) + i d xJ · A , (135) the material, that could affect its optical properties. The first two lines of Eq. (133) are the free Lagrangian where the effective action Se f f (A) has the following explicit densities of these three fields, respectively. form Z   The third line, instead, contains all the relevant interactions, 4 ε0 2 1 2 Se f f (A) = d x A˙ − (∇ × A) with the first term being the traditional dipole interaction term, 2 2µ0 and the second one representing the polarisation-reservoir in- 1 Z + dt dτ d3r g(r)A˙ (t)Γ(t − τ)A˙ (τ), (136) teraction, that models losses in the system. 2 The geometry of the system under investigation is con- where the tensor quantity Γ(r,t −τ) is directly connected with tained in the shape factor g(x), which equals 1 inside the re- the dielectric function of the effective medium the “dressed" gion of interest, and zero otherwise. electromagnetic field is propagating into. An exhaustive dis- To complete the picture, the medium is characterised by a cussion on the physical meaning and origin of Γ(r,t − τ) is resonant frequency ω , static polarisability β(x), mass density 0 given in Ref. 79. (per unit frequency) ρ(x), and spectral coupling f (ω), which If we substitute in the action above the expressions of the accounts for the frequency distribution of the various decay electric and magnetic fields in terms of the vector potential channels of the medium. The dot in Eq. (133), moreover, A, i.e., E = −∂ A and B = ∇ × A, we can see that S (A) represents derivation with respect to time. t e f f is at most quadratic in A and ∂ A, meaning that, de-facto, it Notice, that although (r,t) contains interacting terms, t LHB describes the free evolution of the dressed field. This means, from the point of view of the electromagnetic field, it is still that the path integral in Eq. (135) can be reduced to a form considered as a free Lagrangian density, as no terms propor- similar to Eq. (130). In the present case, however, it is more tional to powers of E or B (or, equivalently, the vector poten- convenient to write the path integral in Fourier space, as a tial A) higher than two is present. function of the frequency ω, rather than time. By doing this We can then substitute Eq. (133) into Eq. (103) to obtain we then obtain the full path integral of the Huttner-Barnett model as  Z  i ←→ 0 Z n Z0(J) = exp dω dRJ(r,ω) G (R,ω)J(r ,ω) , (137) Z(J,JP,JF ) = DADPDF exp iSHB 2 ←→ Z h io where dR = d3rd3r0, R = r − r0, and G (R,ω) is the dyadic + i d4x J · A + J · P + J · F , P F (134) Green’s function of the dressed field in Fourier space, and it is a solution of the following equation of motion for the dressed where J is the source term for the electromagnetic field, field JP the source term for the matter polarisation field, JF the  2   R 4 2 ∂ 2 0 source term for the reservoir field, and SHB = d xLHB is −δµα ∇ + − ω ε(r,ω)δµα Gαν (r − r ,ω) the Huttner-Barnett action. ∂xµ ∂xα 0 Written in this way, Z(J,JP,JF ) allows one to describe, in = µ0δµν δ(r − r ), (138) a fully quantum manner, not only the dynamics of photons ˜ ˜ (through the current J), but also those of polaritons (through where ε(r,ω) = 1+g(r)Γ(r,ω)/ε0 (with Γ being the Fourier transform of Γ appearing in Eq. (136)). the current JP), i.e., quantum excitations of the matter polar- isation field, and to also account for the quantum features of Alternatively, the Green’s function can be also derived from Eq. (137) by using Eq. (131) as the correlation function be- the loss reservoir, through the current JF . However, for the 0 purposes of this tutorial, we shall focus our attention on pho- tween the electromagnetic field at point r and r , i.e., tons only, and therefore we set JP = 0 = JF , leaving J as the 2 [J] 0 δ Z0 only free parameter. In this limit, the path integrals for P Gµν (r − r ,ω) = 0 δJµ (r,ω)δJν (r ,ω) and F(ω) can be performed explicitly, resulting in an effec- J=0 0 tive path integral for the electromagnetic field propagating in ≡ hAµ (r,ω)Aν (r ,ω)i. (139) 27

B. Nonlinear Quantum Electrodynamics The above can also be interpreted as the proba- bility to generate N photons through a N−th order nonlinear In photonics, the description of nonlinear interactions of process. In this case, the probability is given simply by the the electromagnetic field in a medium are often expressed in modulus square of the quantity defined above, i.e., terms of the induced polarisation, which is, in general, ex- P (r ,··· ,r ) = P | (r ,··· ,r )|2 , (145) pressed as a power series expansion upon the electric field101, N 1 N 0 A 1 N i.e. 2 3 n , Π = ε0(χ1E + χ2E + χ3E + ···), where χn is the -th where, again, P0 is a suitable dimensional constant that trans- order susceptibility tensor, containing the linear (n = 1) and forms |A |2 into a probability density. nonlinear (n ≥ 2) response of the medium to the electric field. This form of self-interaction can be generated with an in- teraction Lagrangian similar to that of Eq. (119), i.e., XI. APPLICATIONS OF PATH INTEGRALS IN QUANTUM OPTICS ∞ 1 n Lint (A) = ∑ χn−1(r,ω) · A . (140) n=3 n! We now present two examples on how this formalism can be useful to describe linear and nonlinear quantum optical The lowest order nonlinear effect described by the interac- phenomena in arbitrary media, hoping that this will serve the tion Lagrangian above is that of the interaction of three elec- reader to get a better insight on the potential applications of tromagnetic fields (photons), i.e., a second-order nonlinear path integrals in photonics. 101 process . In this section we then present the solution, by means of We can then write the path integral for the nonlinear in- path integrals, to two well-known problems: (1) the genera- teraction in analogy with Sect. VIII and expand it perturba- tion of spontaneous parametric down conversion (SPDC) in tively in powers of the various χn. However, for the particular a lossy medium, and (2) the calculation of the spontaneous case of nonlinear optics, the magnitude of the various nonlin- emission rate of an atom. Both examples make use of the ear effects gets progressively smaller, as n increases, i.e., that dressed electromagnetic field introduced in Sect. X A, for |χn+1|  |χn| holds for any n. both the interacting and free case, respectively. If we now expand the path integral in powers of χ using Eq. (132) as guidance, we can decompose the path integral in the following sum of three contributions A. Spontaneous Parametric Down Conversion in Lossy Media ∞ (k) Z (J) = Z0(J) + ∑ Z (J) + Zcross(J), (141) k=2 Let us consider only the term n = 3 in Eq. (140). The interaction Lagrangian then becomes (k) where Z0(J) is the free theory given by Eq. (137), Z (J) 1 3 accounts for the k−th order nonlinearity, i.e., it contains terms Lint (A) = χ2(r,ω) · A . (146) proportional to χk solely, and Zcross(J) describes the com- 3! bined interaction of several orders of nonlinearities (for ex- This interaction term, as it is well-known within the nonlin- ample, it contains terms of the form χkχn, with k 6= n etc.). ear optics community, is responsible for second-order nonlin- Since |χn+1|  |χn| holds, however, we can safely neglect ear phenomena, such as second-harmonic generation (SHG), this last term39, as it contains terms that are O(|χ|k). The ex- sum- and difference-frequency generation (SFG/DFG), and plicit expression of the term Z (k)(J) closely resembles that spontaneous parametric down conversion (SPDC)101. As it of the perturbative expansion for an interacting vector field can be seen, Lint involves the interaction of three fields, which theory, defined in Eq. (132), i.e., are conventionally called pump, signal and idler. A very com- n mon assumption (and experimental working condition) is to ∞ "  k# Z χk−1 δ stimulate the aforementioned phenomena by means of a very Z (k)(J) = dω d3r · Z (J)(142). ∑ 1/n 0 bright pump beam, whose quantum properties are typically n=1 (n!) δJ neglected, and whose (i.e., number of photons) does This result can be used to calculate the cross section not change significantly during the interaction106. This ap- (N) σ (r1,··· ,rN) of a given N−th order nonlinear event to take proximation is called undepleted pump approximation. place, as simply the vacuum N-point correlation function of If we accept this approximation, we can rewrite the interac- the electromagnetic field as tion Lagrangian above in a form that allows us to easily cal- culate the path integral. In doing so, essentially, we make use (N) σ (r1,··· ,rN) = σ0 A (r1,··· ,rN), (143) of the undepleted pump approximation to eliminate the pump field from the dynamics (i.e., we consider only the dynam- where σ0 is a suitable dimensional constant and ics of the signal and idler photons), by defining an effective (p) nonlinear susceptibility χ˜µν = χµντ Aτ , so that Eq. (146) A (r1,··· ,rN) = hAµ1 (r1,ω)··· AµN (rN,ω)i becomes39 δ NZ (J) = .(144) [A] = ˜ (r, )A(s)A(i), δJµ1 (r1,ω)···δJµN (rN,ω) J=0 Lint χµν ω µ ν (147) 28

where the superscripts {p,s,i} stand for pump, signal and ���

idler field, respectively (mimicking the usual notation of non- ] Δk=0 linear optics). ��� Substituting the interaction Lagrangian above into Eq. Δk=5 (142) with k = 3 (the order of the nonlinearity in the La- ��� grangian) and n = 1 (first order perturbation term), we can Δk=10 ������� ���� write down the explicit expression of the partition function for [ ��� second-order nonlinear optical processes. In terms of Feyn- man diagrams, the path integral reads

���� ��� � (s) Jµ Gσ µ ��� ��� ��� ��� ��� Z1(J) = (i) + permutations, (148) Gσν Γ �

Jν FIG. 10. Plot of the probability of observing a SPDC event in a 1D homogeneous lossy medium as given by Eq. (154), as a function of where the permutations have to be understood as all the possi- the dimensionless parameter x = ΓL, and for different values of the ble ways to order the three lines (and the correspondent la- phase mismatch parameter, namely ∆k = 0 (blue, solid line), corre- bels) appearing in the Feynman diagram. The dashed line sponding to the case of perfect phase matching, ∆k = 5 (red, solid in the diagram above represents the dressed vacuum state line), and ∆k = 10 (black, solid line). |0i ≡ |αpump;0signal;0idleri, i.e., the effective quantum vacuum seen by the signal and idler photons. From Eq. (148), we can calculate the probability for an material the dyadic Green’s function can be calculated explic- SPDC event to occur using Eqs. (144) and (145) with N = 2, itly using Eq. (111) (with m = 0, k2 = κ2 − ω2 and by using namely the residue theorem to perform the ω−integral) and has the 81 (s) (i) 2 following form PSPDC = hAµ (x,ω)Aν (y,ω)i 1 n G(x − y, ) = (x − y) [i( ( )(x − y) − t)] δ 2Z (J) 2 ω Θ exp κ ω ω = 2iκ(ω) δJ (x,ω)δJ (y,ω) o µ ν + Θ(y − x)exp[i(κ(ω)(x − y) − ωt)] , (152) G signal ( ) = k( ) + i ( ) 2 where κ ω ω γ ω is the complex wave vector of the field in the medium. = G Substituting Eqs. (151) and (152) into Eq. (149) we obtain idler 2  L ( k − i )L Γ ∆ Γ PSPDC = exp − sinc , (153) 2 2 2 (2) = Xµν (x − y) , (149) where Γ = γs +γi are the losses seen by the signal (γs ≡ γ(ωs)) (2) and the idler (γi ≡ γ(ωi)) and ∆k = k(ωp) − k(ωs) − k(ωi) is where Xµν (x−y) is the biphoton propagator, whose explicit expression is given as the mismatch parameter. One can readily see that in the limit Γ = 0, the equation above reproduces the well known result Z 101 2 (2) 4 from nonlinear optics , that PSPDC ∝ sinc (∆kL/2). For the X (x − y) = d z χ˜ (z)Gµα (x − z,ωs)G (z − y,ωi), µν αβ βν general case where Γ 6= 0, instead, the equation above reduces (150) to where ω are the frequencies of the signal and idler fields, s,i 2exp(−ΓL)[cos(∆kL) − cosh(ΓL)] respectively. The probability of detecting an SPDC event is PSPDC = − , (154) then completely determined by the propagators for the signal L2(∆k2 + Γ2) and idler field. As an explicit example of this, let us consider which is depicted in Fig. 10 as a function of the dimensionless a lossy isotropic 1D medium of length L. Since the medium quantity x = ΓL, for different values of the phase mismatch is isotropic, χ2(r,ω) ≡ χ. Moreover, let us assume that the parameter ∆k. pump beam can be modelled as a plane wave of frequency ωp propagating along the x-direction. The effective nonlinear susceptibility then has the following explicit form B. Spontaneous Emission Decay Rate of a Quantum Emitter in an Arbitrary Medium χ˜αβ (x) = χAp exp[i(kpx − ωpt)]. (151) To calculate the biphoton propagator, we need the explicit ex- The spontaneous emission decay rate of a quantum emitter pression for the Green’s function. For a lossy, isotropic, 1D embedded in a homogeneous environment can be written as 29 follows117 In QFT, these loop corrections need to be properly renor- ( ) malised, as they amount to a sort of self-energy of the vacuum 2πcΓ0 n←→ o (in our case self-energy of the dressed electromagnetic field), Γsp = Im Tr G (r,r;ω0) , (155) ω0 which typically diverge. However, while the real part of the Green’s function above actually diverges, its imaginary part where Γ0 is the free-space spontaneous emission decay rate remains bounded, and it is actually the spontaneous emission (the Einstein A coefficient), ω0 is the characteristic frequency rate of the quantum emitter, i.e., of the quantum emitter, here modelled as a two-level system ←→ with a dipole moment µ, and G (r,r;ω) is the dyadic Green Γsp = Im{Zcorr} ∝ Im{Gzz(0,ω0)} (160) function of the electromagnetic field in which the emitter is embedded, calculated at the position of the emitter itself, and To prove that, let us first write down the explicit expres- 81 sion of the loop propagator Gzz(0,ω0), which is now a scalar Tr{A} = ∑Aii is the trace operation . Equation (155) is a i quantity, using Eq. (111) as general result, that holds for different systems, such as atoms in free space117, cavity QED118, and even metamaterials119. Z d3k 1 Gzz(0,ω ) = , (161) If we assume to consider a homogeneous surrounding 0 (2π)3 k2 + m2 medium characterised by the complex-valued dielectric func- 2 2 tion ε = ε1 + iε2, and assume as well that the dipole moment where m = ω0 (ε1 +iε2), i.e., is the “mass of the dressed pho- of the emitter is aligned along the z axis, Eq. (155) simplifies ton in the medium described by the complex dielectric func- to120,121 tion (ε1 + iε2). Since the environment surrounding the quan- tum emitter is homogeneous and isotropic, we can perform √ Γ = Γ0 Re ε1 + iε2 (156) the integral above in spherical coordinates, readily integrat- ing away the angular degrees of freedom, which amount to an We can arrive at the same result of both Eq. (155) and Eq. overall 4π multiplicative term. Separating the real and imagi- (156) using path integrals. To show that, we can use the in- nary part of the fraction in the integral above leads then to the teracting theory of the dressed electromagnetic field described following result in Sect. XI A, but rather than calculating the actual nonlinear ∞ 2 2 2 interaction of the field with itself (i.e., the second-order non- 1 Z k (k + ω ε1) G (0,ω ) = dk 0 linear processes) we turn our attention to the nonlinear cor- zz 0 2 2 2 2 2π 0 [k − ω0 (ε1 + iε2)] rections to the free propagator Z (J). These corrections are 0 iω2ε Z ∞ k2 known in QFT as loop corrections, or, sometimes, vacuum − 0 2 dk . (162) 8,10,88,109 2 2 2 2 bubbles and are represented by Feynman diagrams 2π 0 [k − ω0 (ε1 + iε2)] like the following Both integrals above can be readily evaluated, giving Re{Gzz(0,ω0)} → ∞, and

Zcorr = + higher order terms. (157) ω0 √ Im{Gzz(0,ω )} = Re{ ε + iε }, (163) 0 4π 1 2 This term can be obtained by Eq. (148) by simply identifying which, apart from an inessential multiplicative factor, looks the two source terms Jµ and Jν with each other, and corre- very similar to the result in Eq. (156). sponds to the following integral Z 1 4 Zcorr = − d xλ (x,ω)G (0,ω), (158) XII. CONCLUSIONS AND FUTURE PERSPECTIVES 2 αα αα where ω0 is the transition frequency of the quantum emit- In this tutorial, we have introduced the reader to the con- ter. Notice, that in the integral above, we have used λαα (x), cept and framework of path integrals. We have discussed their rather than χ˜αα (x), since the interaction vertex in this case origin, historical development and their interpretation as sums doesn’t necessarily represent an actual nonlinear optical in- over all possible trajectories, and, in Sect. II, we have pro- teraction, but rather the quantum emitter which, initially its vided the foundations upon which the very concept of path , emits a photon, which gets absorbed at a later integral is rooted. To help the reader familiarise with the for- time (from here the loop term appearing in the Feynman di- malism, and the kind of mathematics needed to deal with path agram above). Moreover, the dipole moment µ of the quan- integrals, in Sect. III we have presented some introductory tum emitter is aligned along the α direction. If, without any standard textbook examples, namely the free particle (Sect. loss of generality, we assume that α → z, and that the quan- III A), the refraction of a light beam from an interface (Sect. tum emitter is localised at the vertex, and that it is charac- III B) , the double slit experiment (Sect. III C), and the quan- terised by a transition frequency ω0, the interaction vertex tum harmonic oscillator (Sect. III D). All these examples have 2 λαα (x,ω) = |µ| δ(ω − ω0) and therefore been discussed extensively, pointing out the procedures and techniques to get to the analytical form of the kernel for those 2 Zcorr = |µ| Gzz(0,ω0). (159) special cases. 30

Starting from Sect. IV, then, we have built upon this knowl- their remarkable ability to treat very complex problems in an edge, and presented a series of less trivial applications of easy and accessible way, could represent a significant step for- path integrals in both classical and quantum optics, with the ward towards this goal. twofold aim of presenting an alternative derivation, based on path integrals, of results known in optics, and helping the reader contextualise this framework in an accustomed envi- ACKNOWLEDGEMENTS ronment. In particular, we have discussed the propagation of light through an inhomogeneous medium in Sect. IV, while in The authors acknowledge the financial support of the Sect. V, we have shown how this framework can also be em- Academy of Finland Flagship Programme (PREIN - decisions ployed to solve problems of quantum nonlinear optics, and we 320165, 320166) have discussed explicitly the case of parametric amplification. The first part of our tutorial concludes with Sect. VI, where we give the reader an idea of the computational capabilities of DATA AVAILABILITY PIMC, and briefly discuss how this could be used in photon- ics. Within the current photonics ecosystem, and considering The data that support the findings of this study are available the present challenges faced by this discipline, in fact, we be- from the corresponding author upon reasonable request. lieve that having at disposal a computational framework that allows calculations of complex light-matter interaction sce- narios, and the derivation of optical properties of complex REFERENCES materials in an exact way, without the need to resort to any particular regime of approximation, could represent a very vi- able resource for photonics. This, somehow, is the very aim 1R. P. Feynman, “Space-Time Approach to Non-Relativistic Quantum Me- of our tutorial, i.e., to introduce the photonics community to chanics,” Rev. Mod. Phys. 20, 367–387 (1948). 2 this elegant, but also very powerful, formalism, and stimu- R. P. Feynman and A. R. Hibbs, Quantum Mechanics and Path Integrals (McGraw-Hill, 1965). late cross-contamination between these two seemingly differ- 3J. Zinn-Justin, Path Integrals in Quantum Mechanics (Oxford University ent disciplines. Press, 2005). The second part of our tutorial, starting with Sect. VII gives 4L. M. B. (ed.), Feynman’s Thesis – A New Approach to the reader a very basic introduction on the use of path integrals (World Scientific, 2005). 5H. Kleinert, Path Integrals in Quantum Mechanics, Statistics, Polymer in QFT. Although at a first glance this might seem too ad- Physics, and Financial Markets (World Scientific, 2007). vanced for the spirit of this tutorial, we nevertheless included 6L. S. Schulman, Techniques and applications of path integration (Dover, this topic, for essentially two reasons. First, since many of 2005). the modern problems in theoretical physics, such as quan- 7L. C. Fai, : Feynman Path Integrals and Diagram- tum gravity, , topological field theories, and so matic Techniques in Condensed Matter (CRC Press, 2019). 8A. Das, Field Theory: a Path Integral Approach (World Scientific, 2006). on, are formulated in terms of path integrals, introducing the 9R. J. Rivers, Path Integrals in Quantum Field Theory (Cambridge Univer- reader to the very basic formalism of path integrals in QFT sity Press, 1988). would greatly help them, were they interested in reading more 10M. Maggiore, A Modern Introduction to Quantum Field Theory (Oxford about these topics. Secondly, sections VII-IX set the stage for University Press, 2005). 11L. D. Fadeev and A. A. Slavnov, Gauge Fields: An Introduction to Quan- Sect. X, where we present a unified way to describe quantum tum Theory, 2nd ed. (Westview, 1991). nonlinear dynamics of the electromagnetic field in arbitrarily 12P. H. Frampton, Gauge Field Theories (Wiley, 2008). shaped media in terms of path integrals, as an example of the 13S. W. Hawking, “ and path integrals,” Phys. Rev. D 18, potential that this framework has to offer to photonics, not 1747 (1978). 14 only from a purely computational perspective, like the predic- N. D. Birrell and P. C. W. Davies, Quantum Fields in Curved Space (Cam- bridge University Press, 1982). tion of optical properties of materials granted by PIMC, but 15J. Polchinski, String Theory. Volume 1. An Introduction to the Bosonic also as a viable and simple analytical tool for complex prob- String (Cambridge University Press, 2005). lems in photonics. 16J. M. Mourao, “Aspects of the connections between path integrals, quan- With this tutorial, we hope to have sparked the curiosity tum field theory, topology and geometry,” in Proceedings of the XII Fall Workshop on Geometry and Physics, Coimbra 2003 (Spanish Royal Math- and interest of the reader on the topic, and conveyed the mes- ematical Society, Coimbra, 2004). sage that path integrals do not only represent an elegant for- 17E. L. Pollock and D. M. Ceperley, “Simulation of quantum many-body mulation of quantum mechanics , but they also provide the systems by path-integral methods,” Phys. Rev. B 30, 2555–2568 (1984). 18D. M. Ceperley, R. O. Simmons, and R. C. Blasdell, “Kinetic energy of language of modern theoretical physics, and can be applied in 4 many different areas of science. For the specific case of pho- liquid and solid He,” Phys. Rev. Lett. 77, 115–118 (1996). 19D. M. Ceperley, “Path integrals in the theory of condensed helium,” Rev. tonics, as extensively discussed in this tutorial, path integrals Mod. Phys. 67, 279–355 (1995). could represent a valuable analytical and computational re- 20D. M. Ceperley, “Path integral monte carlo methods for fermions,” in source for investigating complex and multi-physics photonics Monte Carlo and Molecular Dynamics of Condensed Matter Systems, devices, which would require the assembly of model systems edited by K. Binder and G. Ciccotti (Editrice Compositori, Bologna, 1996). from different disciplines, a task that is not always simple to 21C. Pierleoni, D. M. Ceperley, B. Bernu, and W. R. Magro, “Equation carry out. The universality of path integrals, the simplicity of state of the hydrogen plasma by path integral monte carlo simulation,” of adaption of the formalism to very different situations, and Phys. Rev. Lett. 73, 2145–2149 (1994). 31

22B. Militzer and D. M. Ceperley, “Path integral monte carlo simulation of 49M. Ammon and J. Erdmenger, Gauge/Gravity Duality. Foundations and the low-density hydrogen plasma,” Phys. Rev. E 63, 066404 (2001). Applications (Cambridge University Press, 2015). 23V. Gorelov, M. Holzmann, D. M. Ceperley, and C. Pierleoni, “Energy gap 50J. Baez and J. P. Muniain, Gauge Fields, Knots and Gravity (World Scien- closure of crystalline molecular hydrogen with pressure,” Phys. Rev. Lett. tific, 1994). 124, 116401 (2020). 51M. B. Green, J. H. Schwarz, and E. Witten, (Cam- 24R. P. Feynman, Statistical Mechanics: A Set of Lectures (CRC Press, bridge University Press, 2012). 1998). 52J. Tempere and J. P. Devreese, Path-Integral Description of Cooper Pair- 25K. V. Bhagwat, D. C. Khandekar, and S. V. Lawande, Path Integral Meth- ing, in Superconductors – Materials, Properties and Applications, Alexan- ods and their Applications (World Scientific, 1993). der Gabovich (IntechOpen, 2012). 26A. V. Reznichenko, I. S. Terekhov, and S. K. Turitsyn, “Calculation of 53I. H. Duru and H. Kleinert, “Quantum mechanics of H-atom mutual information for nonlinear optical fiber communication channel at from path integrals,” Fortschritte der Physik 30, 401–435 (1982), large SNR within path-integral formalism,” Journal of Physics: Confer- https://onlinelibrary.wiley.com/doi/pdf/10.1002/prop.19820300802. ence Series 826, 012026 (2017). 54M. Creutz and B. Freedman, “A statistical approach to quantum mechan- 27K. Zaheer and M. S. Zubairy, “Atom-field interaction without the rotating- ics,” Ann. Phys. 132, 427–462 (1981). wave approximation: A path-integral approach,” Phys. Rev. A 37, 1628– 55C. Morningstar, “The Monte Carlo method in quantum field theory,” e- 1633 (1988). print arXiv:hep-lat/0702020 (2007). 28V. Chernyak and S. Mukamel, “Path integral formulation of retardation 56M. Creutz, Quarks, and Lattices (Cambridge University Press, effects in nonlinear optics,” The Journal of 100, 2953– 1983). 2974 (1994), https://doi.org/10.1063/1.466438. 57C. Gattringer and C. B. Lang, on the Lattice: 29Y. Tanimura and S. Mukamel, “Real-time path-integral approach to quan- An Introductory Presentation (Springer, 2010). tum and dephasing in nonadiabatic transitions and nonlinear 58M. J. E. Westbroek, P. R. King, D. D. Vvedensky, and S. Dürr, “User’s optical response,” Phys. Rev. E 47, 118–136 (1993). guide to Monte Carlo methods for evaluating path integrals,” Am. J. Phys. 30L. T. Perelman, J. Wu, I. Itzkan, and M. S. Feld, “Photon migration in 86, 293–304 (2018). turbid media using path integrals,” Phys. Rev. Lett. 72, 1341–1344 (1994). 59R. C. Blasdell, D. M. Ceperley, and R. O. Simmons, “Neutron and pimc 31C. Gómez-Reino and J. Liñares, “Optical path integrals in gradient-index determinations of the longitudinal momentum distribution of hcp, bcc and media,” Journal of the Optical Society of America A 4, 1337 (1987). normal liquid 4he,” Z. Naturforsh 48A, 433 (1993). 32M. C. Braidotti, C. Conti, M. Faizal, S. Dey, L. Alasfar, S. Alsaleh, and 60C. PIerleoni, W. R. Magro, D. M. Ceperley, and B. Bernu, “Path inte- A. Ashour, “Path integral for non-paraxial optics,” EPL (Europhysics Let- gral monte carlo simulation of hydrogen plasma,” in Proceedings of the ters) 124, 44001 (2018). (Binz Germany) International Conference on the Physics of Strongly Cou- 33Y. Dimant and S. Levit, “Path integrals for light propagation in dielectric pled Plasmas, edited by W. D. Kraeft and M. Schlanges (World Scientific, media,” J. Opt. Soc. Am. B 27, 899–903 (2010). Singapore, 1996). 34J. I. Gersten and A. Nitzan, “Path-integral approach to electromagnetic 61M. Leino, J. Nieminen, and T. T. Rantala, “Finite temperature quantum phenomena in inhomogeneous systems,” J. Opt. Soc. Am. B 4, 293–298 distribution of hydrogen adsorbate on nickel (001) surface,” Surface Sci- (1987). ence 600, 1860–1869 (2006). 35E. M. Wright and J. C. Garrison, “Path-integral derivation of the complex 62M. Leino, I. Kylänpää, and T. T. Rantala, “Coverage dependence of finite abcd huygens integral,” J. Opt. Soc. Am. A; (United States) 4:9 (1987), temperature quantum distribution of hydrogen on nickel(001) surface,” 10.1364/JOSAA.4.001751. Surface Science 601, 1246–1254 (2007). 36S. Vatsya, “Path-integral formulation of optical beam propagation,” J. Opt. 63I. Kylänpää, M. Leino, and T. T. Rantala, “Hydrogen molecule ion: Path- Soc. Am. B 22, 2512–2518 (2005). integral monte carlo approach,” Phys. Rev. A 76, 052508 (2007). 37M. Hillery and M. S. Zubairy, “Path-integral approach to the quantum the- 64I. Kylänpää, T. T. Rantala, and D. M. Ceperley, “Few-body reference data ory of the degenerate parametric amplifier,” Phys. Rev. A 29, 1275–1287 for multicomponent formalisms: Light-nuclei molecules,” Phys. Rev. A (1984). 86, 052506 (2012). 38M. Hillery and M. S. Zubairy, “Path-integral approach to problems in 65J. Tiihonen, I. Kylänpää, and T. T. Rantala, “Adiabatic and nonadiabatic quantum optics,” Phys. Rev. A 26, 451–460 (1982). static polarizabilities of h and h2,” Phys. Rev. A 91, 062503 (2015). 39M. Difallah, A. Szameit, and M. Ornigotti, “Path-integral description of 66J. Tiihonen, I. Kylänpää, and T. T. Rantala, “Static field- quantum nonlinear optics in arbitrary media,” Phys. Rev. A 100, 053845 gradient polarizabilities of small atoms and molecules at finite tem- (2019). perature,” The Journal of Chemical Physics 147, 204101 (2017), 40B. Cox and J. R. Forshaw, The Quantum Universe (Penguin, 2012). https://doi.org/10.1063/1.4999840. 41N. Wiener, “The average value of a functional,” Proc. London Math. Soc. 67I. Kylänpää and T. T. Rantala, “First-principles simulation of molec- 22, 454–467 (1924). ular dissociation–recombination equilibrium,” The Journal of Chemical 42S. G. Brush, “Functional Integrals and Statistical Physics,” Rev. Mod. Physics 135, 104310 (2011), https://doi.org/10.1063/1.3633516. Phys. 33, 79–92 (1961). 68M. Leino and T. T. Rantala, “Temperature effects on electron correlations 43M. Chaichian and A. Demichev, Path Integrals in Physics: Volume I, in two coupled quantum dots,” Few-Body Systems 40, 237 (2007). Stochastic Processes and Quantum Mechanics (CRC Press, 2001). 69M. Leino and T. T. Rantala, “Finite-temperature effects on correlation of 44G. Bacciagaluppi and A. Valentini, Quantum Theory at the Crossroads: electrons in quantum dots,” Physica Scripta T114, 44–48 (2004). Reconsidering the 1927 Solvay Conference (Cambridge University Press, 70J. Tiihonen, A. Schramm, I. Kylänpää, and T. T. Rantala, “Exact modeling 2009). of finite temperature and quantum delocalization effects on reliability of 45P. A. M. Dirac, “The Lagrangian in quantum mechanics,” Phys. Z. Sowje- quantum-dot cellular automata,” Journal of Physics D: tunion 3, 64–72 (1933). 49, 065103 (2016). 46P. A. M. Dirac, The Principles of Quantum Mechanics (Oxford University 71I. Kylänpää and T. T. Rantala, “Thermal dissociation of dipositronium: Press, 1947). Path-integral monte carlo approach,” Phys. Rev. A 80, 024504 (2009). 47It is not difficult, using an electromagnetic analogy, to see that in this con- 72I. Kylänpää and T. T. Rantala, “Finite temperature quantum statistics of text, Planck’s constant h plays the role of wavelength of the path. A plane h3+ molecular ion,” The Journal of Chemical Physics 133, 044312 (2010), wave propagating, say, along the x-direction, in fact, is associated with a https://doi.org/10.1063/1.3464758. phase factor of the form exp(i2π x/λ). The phase factor carried by each 73J. Tiihonen, I. Kylänpää, and T. T. Rantala, “Computation of dynamic path i, instead, goven by exp(i2π S/h). A simple comparison allows the polarizabilities and van der waals coefficients from path-integral monte reader to identify λ → h. carlo,” Journal of Chemical Theory and Computation 14, 5750–5763 48S. Weinberg, The Quantum Theory of Fields - Volume 2: Modern Applica- (2018), pMID: 30278124, https://doi.org/10.1021/acs.jctc.8b00859. tions (Cambridge University Press, 2005). 74J. Tiihonen, I. Kylänpää, and T. T. Rantala, “General polarizability and hy- perpolarizability estimators for the path-integral monte carlo method ap- 32

plied to small atoms, ions, and molecules at finite temperatures,” Phys. 105R. J. Glauber, “Coherent and Incoherent States of the Radiation Field,” Rev. A 94, 032515 (2016). Phys. Rev. 131, 2766–2788 (1963). 75I. Ruokosenmäki and T. T. Rantala, “Numerical path integral approach 106P. D. Drummond and M. Hillery, The Quantum Theory of Nonlinear Optics to quantum dynamics and stationary quantum states,” Communications in (Cambridge University Press, 2014). 18, 91–103 (2015). 107J. Tiihonen, Thermal Effects in Atomic and Molecular Polarisabilities with 76I. Ruokosenmäki, H. Gholizade, I. Kylänpää, and T. T. Rantala, “Numeri- Path Integral Monte Carlo (Tampere University Dissertations 35, 2019). cal path integral solution to strong coulomb correlation in one dimensional 108A. Anton, J. Henri, D. Yunyun, W. Yadong, L. Harri, and S. Zhipei, “Non- hooke’s atom,” Computer Physics Communications 210, 45–53 (2017). linear optics with 2d layered materials,” Advanced Materials 30, 1705963 77H. Gholizadehkalkhoran, I. Ruokosenmäki, and T. T. Rantala, “Eigen- (2018), https://onlinelibrary.wiley.com/doi/pdf/10.1002/adma.201705963. states and dynamics of hooke’s atom: Exact results and path inte- 109M. Srednicki, Quantum Field Theory (Cambridge University Press, 2007). gral simulations,” Journal of 59, 052104 (2018), 110D. M. Bishop and J. Pipin, “Dipole, quadrupole, octupole, and https://doi.org/10.1063/1.5028503. dipole–octupole polarizabilities at real and imaginary frequencies for h, 78I. Ruokosenmäki and T. T. Rantala, “Real-time diffusion monte carlo he, and h2 and the dispersion-energy coefficients for interactions between method,” Communications in Computational Physics 25, 347–360 (2018). them,” International Journal of 45, 349–361 (1993), 79A. Bechler, “Quantum electrodynamics of the dispersive dielectric https://onlinelibrary.wiley.com/doi/pdf/10.1002/qua.560450403. medium–a path integral approach,” J. Mod. Opt. 46, 901–921 (1999). 111Y. Kawamura, “Fermionic scalar field,” (2014), arXiv:1406.6155 [hep-th]. 80We implicitly assumed that the two set of coordinates span a one dimen- 112A scalar electromagnetic field is defined as an electromagnetic wave whose sional space, for simplicity. Extension to higher dimensions can be, in fact, polarisation is the same in any point of its wavefront, and always remains easily carried out. the same during propagation. If the interaction of a scalar electromagnetic 81F. W. Byron and R. W. Fuller, Mathematics of Classical and Quantum field with a system causes its state of polarisation to vary, this variation will Physics (Dover, 1992). happen in the same manner in every point of its wavefront. A good exam- 82J. J. Sakurai, Modern Quantum Mechanics (Revised Edition) (Addison- ple of a scalar electromagnetic field is an optical beam, which is typically Wesley, 1994). written as E(r,t) = ψ(r,t)uˆ, where uˆ is a constant (unitary) transverse 83V. I. Arnold, Mathematical Methods of (Springer, vector (with respect to the rpopagation direction of the field), which de- 1997). scribes the field polarisation. The dynamics of an optical beams are then 84W. Heisenberg, “über quantentheoretische umdeutung kinematischer und fully determined by ψ(r,t), which, in general, is a solution of the scalar mechanischer beziehungen,” Z. Phys. 33, 879–893 (1925). . 85M. Born and P. Jordan, “Zur quantenmechanik,” Z. Phys. 34, 858–888 113E. A. Calzetta and B.-L. B. Hu, Nonequilibrium Quantum Field Theory (1925). (Cambridge University Press, 2008). 86T. F. Jordan, Quantum Mechanics in Simple Matrix Form (Dover, 2006). 114M. Creutz, “Quantum electrodynamics in the temporal gauge,” Annals of 87E. Schrödinger, “Quantisierung als eigenwertproblem,” Ann. Phys. 384, Physics 117, 471–483 (1979). 361 (1926). 115W. Vogel and D. G. Welsch, Quantum Optics, 3rd ed. (Wiley, 2006). 88L. S. Brown, Quantum Field Theory (Cambridge University Press, 1994). 116B. Huttner and S. M. Barnett, “ of the electromagnetic field 89M. Kaku, Quantum Field Theory: a Modern Introduction (Oxford Univer- in dielectrics,” Phys. Rev. A 46, 4306–4322 (1992). sity Press, 1993). 117R. Loudon, The Quantum Theory of Light (Oxford Science, 2000). 90L. D. Landau and E. M. Lifchitz, Classical Theory of Fields (Butterworth 118H. Walther, B. T. H. Varcoe, B.-G. Englert, and T. Becker, “Cavity quan- Heinemann, 1996). tum electrodynamics,” Reports on Progress in Physics 69, 1325–1382 91L. Fetter and J. D. Walecka, Quantum Theory of Many-Particle Systems (2006). (Dover, 2003). 119M. Lobet, I. Liberal, E. N. Knall, M. Z. Alam, O. Reshef, R. W. Boyd, 92R. K. Luneburg, Mathematical Theory of Optics (Cambridge University N. Engheta, and E. Mazur, “Fundamental radiative processes in near-zero- Press, 1965). index media of various dimensionalities,” ACS Photonics 7, 1965–1970 93R. P. Feynman, QED - The Strange Theory of Light and Matter (Penguin (2020). Press Science, 1990). 120S. M. Barnett, B. Huttner, and R. Loudon, “Spontaneous emission in ab- 94R. P. Feynman, The Feynman Lectures on Physics - Vol. 3 (Basic Books sorbing dielectric media,” Phys. Rev. Lett. 68, 3698–3701 (1992). (New Millenium Edition), 2010). 121C.-A. Guérin, B. Gralak, and A. Tip, “Singularity of the dyadic green’s 95J. D. Jackson, Classical Electrodynamics (Wiley, 1986). function for heterogeneous dielectrics,” Phys. Rev. E 75, 056601 (2007). 96Notice that the pre-factor of Eq. (62) differs slightly from Eq. (25) of Ref. 31. This is due to the different dimensionality of the problem at hand. While Ref. 31 deals with a (2 + 1)-dimensional problem, in our model we are only considering√ one transverse dimension, and therefore the pre-factor has the form of X instead of X. Appendix A: How to Deal with Path Integrals in QFT: 97K. D. C. and S. V. Lawande, “Exact propagator for a time Reducing Eq. (134) to Eq. (135) Using Path Integration dependent harmonic oscillator with and without a singular per- turbation,” Journal of Mathematical Physics 16, 384–388 (1975), https://aip.scitation.org/doi/pdf/10.1063/1.522511. In this appendix, we show how one can perform the path 98F. W. J. Olver, D. W. Lozier, R. F. Boisvert, and C. W. Clark, NIST Hand- integration over the matter P and reservoir F(ω) fields in Eq. book of Mathematical Functions (Cambridge University Press, 2010). (134), converting the full path integral partition function to the 99O. Steuernagel, “Equivalence between focused paraxial beams and the effective partition function Ze f f (J) shown in Eq. (135). This quantum harmonic oscillator,” American Journal of Physics 73, 625 (2005). will introduce to the reader not accustomed to this field how 100M. Fox, Quantum Optics: An Introduction (Oxford University Press, one can perform explicit calculations relating to path integra- 2006). tion for fields39,79. 101 R. W. Boyd, Nonlinear Optics (Academic Press, 2008). Before we begin our calculations, it is useful to recall 102T. Lancaster and S. J. Blundell, Quantum Field Theory for the Gifted Am- ateur (Oxford University Press, 2014). the explicit formula for calculating a Gaussian integral in 7,102 103C. C. Gerry and P. L. Knight, Introductory Quantum Optics (Cambridge n dimensions , which will be useful for the calculations University Press, 2005). presented in this appendix. To do so, we first need to de- 104E. C. G. Sudarshan, “Equivalence of semiclassical and quantum mechan- n n fine two vectors x = (x1,x2,...,xn) ∈ R , and b ∈ R , a non- ical descriptions of statistical light beams,” Phys. Rev. Lett. 10, 277–279 singular n × n matrix A ∈ C, such that the scalar product (1963). T (x,Ax) = xi Ai jx j is a quadratic form, the following result 33 holds: and r Z ∞ Z  1  πn 1  ˙ dn x exp − (x,Ax) − (b,x) = exp (b,A−1b) , Lmr (P,F) = −g(x) dω f (ω,x)P · F. (A9) 2 detA 2 0 In the above equations, a dot indicates derivative with respect −1 where A is the inverse of A. to time. This result can be used to compute Gaussian integrals over Notice, how Eq. (A4) only contains quantities depending field configurations. In analogy with what has been done in on the reservoir field F(ω) (with the matter field P acting as Sect. III for quantum particles, one can in fact discretise the a parameter, with respect to DF(ω)), and no term of order path integral over the field configuration into a sum over a fi- higher than F2(ω) appears, meaning that we can compute this nite set of configurations, i.e., dP ∝ ∏k dP(xk), perform Gaus- integral by means of Gaussian integration. sian integration with respect to xk, and then take the usual con- tinuum limit. We begin by restating the full path integral of the Huttner- 1. Integration over the reservoir fields 116 Barnett model as given in Sec. X, setting JP = JF = 0: Z  Z  Let us start by rewriting the exponent in Eq. (A4) in a form 4 ˆ Z(J) = DADPDF exp iSHB + i d xJ · A . (A1) that will explicitly contain a quadratic form (F(ω),AF(ω)) in the reservoir field. We then first integrate by parts (with respect to time), use the identity As we have set JP = JF = 0 (because, as said in the main text, we are only interested in photon dynamics, and not in the 2 ∂φ  ∂  ∂φ  ∂ 2φ dynamics of polaritons or other quasi-particles), the only free = φ − φ 2 , (A10) parameter is the source term J for the electromagnetic field. ∂t ∂t ∂t ∂t Moreover, since we want to reduce the path integral above to and integrate again by parts, to get the following result: Eq. (135), we can first factor out the parts of the Huttner- Z Z ∞ Barnett Lagrangian that depend only on the electromagnetic 0 3 n Sres (F) + Smr (F,P) = dt dt d x dω g(x) field, including the J·A-term above. To do so, we group them 0 into the term S (A), in order to rewrite the integral above in 1 o em − F(t0), Aˆ(t,t0)F(t) − b(t0),F(t) ,(A11) the following, nested form 2 Z where x is relabelled as x from now on in this appendix, for Z(J) = DA exp[iSem (A)] IP (A), (A2) simplicity, and we have defined

0 0 with b(t ) = −ig(x) f (ω)δ(t −t )P˙ , (A12) Z   and the operator Aˆ ≡ Aˆ(t,t0) as IP (A) = DP exp i Smat (P) + Sm f (A,P) IF (P), (A3)  2  0 iρg(x) ∂ 2 0 and Aˆ(t,t ) = + ω δ(t −t ). (A13) 2 ∂t2 Z IF (P) = DF exp[i(Sres (F) + Smr (F,P))], (A4) Notice, how we have added a second time integration (and consequently a Dirac delta function δ(t − t0)) in Eq. (A11) where the various actions Sk appearing above sum to the total to facilitate the definition of the quantities presented above. R 4 ˆ 0 Huttner-Barnett action SHB = d xLHB, where the Huttner- Notice, moreover, how the operator A(t −t ) is very similar to Barnett Lagrangian is given in Eq. (133), from which the fol- the usual wave equation operator109,117. lowing individual pieces can be easily defined: It is instructive to repeat the salient steps that led to Eq. (A11). First, integration by parts with respect to time allows ε0 2 1 2 us to rewrite the reservoir-matter field interaction term as Lem (A) = A˙ − (∇ × A) , (A5) 2 2µ0 Z Z dt −g(x) f (ω)P · F˙  = g(x) f (ω) dtP˙ · F. (A14)

g(x) h 2 2 2i Lmat (P) = P˙ − ω P , (A6) Then, the identity presented in Eq. (A10) can be used to 2 0 2 2ε0 ω0 β(x) rewrite the term proportional to F˙ as

ρ g(x) Z ∂F ∂F Z ∞  2  dt dt0 δ(t −t0) , (A15) ρ(x) 2 ρ(x)ω 2 0 Lres (F) = g(x) dω F˙ − F , (A7) 2 ∂t ∂t 0 2 2 where the Dirac delta function δ(t − t0) has been inserted to differentiate the two time derivative terms, thus making the L f m (A,P) = −g(x)A˙ · P, (A8) effect of part integration appear more obvious. At this point, if 34 we integrate by parts, we can shift one of the time derivatives with NF a suitable normalisation constant and in one term onto the other one, thus obtaining hZ Z ∞ | f (ω)|2 g(x) b, Aˆ−1b = −i dt d3x dω P2(t) ρg(x) Z ∂ 2F 0 ρ − dt dt0 δ(t −t0)F . (A16) 2 t2 Z g(x) i ∂ + dt dt0 d3x P(t)G(t −t0,x)P(t0) (A23), ρ Putting everything together, we can easily reconstruct both the linear and the quadratic form in Eq. (A11). where Z ∞ 0 2 2 0 G(t −t ,x) = dω ω | f (ω)| DF (t −t ,ω) (A24) 2. How to calculate the operators Aˆ and Aˆ−1 0 is defined to be the time-domain Green’s function of the reser- The Green’s function for the operator Aˆ in Eq. (A13) can voir field. be found in the following manner. The operator contains no spatial derivatives and so the Green’s function K(t −t0,x −x0) associated with it will have a localized spatial dependence of 4. Integration over the matter fields the form δ(x − x0). Hence, Eq. (A22) is the solution to the integral over the reservoir  ∂ 2  degrees of freedom. As it can be seen, this result contains iρg(x) + ω2 K(t −t0,x) = δ(t −t0). (A17) ∂t2 a dependence on the matter polarisation field P. If we wish to solve the integral IP with the same strategy adopted above Taking the Fourier transform of (A17) with respect to time for IF , i.e., by employing Gaussian integration, we first need produces to rewrite it in a way containing a quadratic form of the kind (P, Bˆ P). To do that, we can first integrate by parts with re- 0 0 2 2 exp(iΩt ) spect to time t and to time t in Eq. (A23), which allows us to −Ω + ω K(Ω,x) = . (A18) 0 iρ g(x) rewrite the term containing P˙ DF (t −t ,x)P˙ as Z 2 0 From these results, one finds the expression for the Green’s 0 00 ∂ DF (t −t ,ω) 0 0 dt dt P(t) P(t ). (A25) function K(t − t ,x) by simply integrating the result above ∂t∂t0 with respect to the frequency Ω, obtaining As the Feynman propagator depends on the difference t − t0, 1 Z dΩ exp[iΩ(t −t0)] one can change variables to τ = t − t0, yielding ∂ 2/∂t∂t0 = K(t −t0,x) = iρg(x) 2π (ω2 − Ω2) ∂ 2/∂τ2. The fact that the Feynman propagator obeys 0 DF (t −t ,ω)  2  ≡ , (A19) ∂ 2 iρg(x) + ω DF (τ,ω) = δ(τ), (A26) ∂τ2 D (t − t0, ) with F ω the Feynman propagator of the reservoir allows us to write field, explicitly given by 2 ∂ DF (τ,ω) 2 Z 0 = δ(τ) − ω DF (τ,ω). (A27) 0 dΩ exp [iΩ(t −t )] 2 DF (t −t ,ω) = . (A20) ∂τ 2π (ω2 − Ω2) Converting back to the separate time variables, t and t0, and Hence, by recalling that the Green’s function of an operator Aˆ inserting our result into Eq. (A25), we get can be interpreted as the inverse of that same operator81, we Z Z 2 2 0 0 0 can make the following formal identification dtP (t) − ω dtdt P(t)DF (t −t ,ω)P(t ). (A28)

0 DF (t −t ,ω) Before proceeding with our calculations, it is useful to intro- Aˆ−1 → K(t −t0,x) = . (A21) iρg(x) duce the following quantity Z ∞ 0 2 2 0 G(t −t ,x) = dωω | f (ω,x)| DF (t −t ,ω), (A29) 3. Calculation of the integral IF 0 which can be interpreted as the reservoir-weighted Green’s If we now discretise the integration measure, i.e., DF → function. With this at hand, the exponent in Eq. (A23) be- ∏k dF(xk,ω) we can perform the integral appearing in Eq. comes (A4) using first Gaussian integration and then taking the con- i hZ Z ∞ | f (ω)|2g(x) tinuum limit. By doing so, we obtain the following result − dt d3x dω P2(t) 2 0 ρ   Z 1 −1  0 3 g(x) 0 0 i IF (P) = NF exp b, Aˆ b , (A22) + dt dt d x P(t)G(t −t ,x)P(t ) . (A30) 2 ρ 35

5. Derivation of the quadratic form for IP complicated (we discuss it briefly in the next subsection, but the interested reader is referred to Ref. 79 for more details). ( ) In the form above, the exponent of IF (P) contains a term Once we have transformed the exponent Φ P into a proportional to P2. This term can be summed with the cor- quadratic form, we can again discretise the path integral, solve responding quadratic term appearing in the free part of the using Gaussian integration, and then take the limit of it, to ob- matter action Smat (P) in the expression of IP(A) given by Eq. tain the following result (A3), i.e.,  Z  i 0 3 ˙ 0 ˙ 0 ( ) IP (A) = exp dt dt d xg(x)A(t)Γ(t −t ,x)A(t ) , Z g(x) Z ∞ | f (ω)|2g(x) 2 dt d3x − − dω P2(t).(A31) 2ε β 2ρ 0 0 where Γ(t −t0,x) is the Green’s function of the operator Bˆ. q 2 If we now introduce the quantity v(ω) = f (ω) ε0ω0 βρ and the scaled resonance frequency 6. Expression and physical meaning of Γ(t −t0,x)

Z ∞ |v(ω)|2 ˜ 2 2 The Green’s function for the operator Bˆ just defined can ω0 = ω0 + dω 2 , (A32) 0 ρ be found in the following way: first, notice that the prefactor ig(x) in the definition of Bˆ, i.e., Eq. (A37) can be treated, for we can rewrite the term in the curly brackets above in the fol- the purpose of finding the Green’s function, as a “numerical" lowing compact form constant, and therefore it can be neglected during calculation (we can put it back at the end of it). Using the definition of g(x) Z ∞ | f (ω)|2 g(x) g(x)ω˜ 2 0 Green’s function81 we then know that Γ(t −t0,x) is a solution − − dω = − 2 . (A33) 2ε0 β 0 2ρ 2ε0 ω0 β of the equation Bˆ Γ(t −t0,x) = δ(t −t0). It should be noted that in this case the operator Bˆ is not Notice, how the primary effect of the reservoir field is to intro- simply a differential operator, as was the case earlier in (A13), duce a frequency-dependent shift in the resonance frequency but it contains an extra source term. The Green’s function ω of the material. 0 for Bˆ is then found as the solution to the following integro- The I -integral now reads P differential equation Z I = P exp[iΦ(P)], (A34) 1  ∂ 2  P D ˜ 2 0 2 2 + ω0 Γ(t −t ,x) ε0 ω0 β ∂t where now 1 Z − dτ G(t − τ,x)Γ(τ −t,x) = δ(t −t0). (A38) Z  g(x)    ρ 3 ˙ 2 ˜ 2 2 ˙ Φ(P) = dt d x 2 P − ω0 P + g(x)A · P 2ε0 ω0 β We can solve the equation above by means of Fourier trans- 1 Z g(x) formation with respect to t −t0, thus obtaining + dt dt0 d3x P(t)G(t −t0,x)P(t0). (A35) 2 ρ 1 1 ˜ 2 2 ˜ ˜ ˜ 2 ω0 − Ω Γ(Ω,x) − G(Ω,x)Γ(Ω,x) = 1, (A39) As we did for IF , one can then integrate by parts the term ε0ω β ρ 2 0 proportional to P˙ and introduce an extra integral with respect to time, to be able to write the above quantity as a quadratic where G˜(Ω,x) is the Fourier transform of G(t − t0,x). After form, i.e., (P,Bˆ P) + (c,P), where a lengthy but straightforward calculation79, one can arrive at the following form for the Fourier transform of Γ(t −t0,x) c(t) = −ig(x)A˙ , (A36) 2 ε0ω0 β 0 Γ˜ (Ω,x) = , (A40) and the expression for the operator Bˆ ≡ Bˆ(t,t ) is obtained by 2 2 ω − Ω [1 + λF (Ω)] looking at the terms sandwiched between the terms P(t0) and 0 P(t), i.e., where λF (Ω) is the Fourier transform of the Feynman propa- 0 gator DF (t −t ,ω). ig(x)  ∂ 2  ˆ 0 ˜ 2 0 The above result can be interpreted as an effective dielectric B(t,t ) = 2 2 + ω0 δ(t −t ) ε0 ω0 β ∂t constant for the “dressed" electromagnetic field. To see this, ig(x) let us consider the full expression of the effective Lagrangian − G(t −t0,x). (A37) ρ coming from Lem as given in Eq. (A5) and the result given by IP(A). Combining these two terms and remembering that Notice how the form of the operator Bˆ is very similar to that of A˙ = −E and ∇ × A = B, we get, in Fourier space the operator Aˆ introduced when calculating IF . The extra term 0 2 1 2 appearing above, i.e., −ig(x)G(t − t ,x)/ρ, however, makes Le f f (E,B) = ε0|E(Ω,x)| + |B(Ω,x)| Bˆ intrinsically integro-differential, which, in turns, makes µ0 the determination of its correspondent Green’s function quite + g(x)Ω2 Γ˜ (Ω,x)|E(Ω,x)|2, (A41) 36 where the effective Lagrangian has been written in terms of the reservoir field), and the possible interaction channels that the electric and magnetic field, since this form will provide an the medium has with the electromagnetic field (through the easier understanding of the interpretation of Γ(t −t0,x). From matter porlarisation). After the integration with respect to F Le f f we can calculate the displacement vector D(Ω,x), as and P the partition function has the following form follows90

∂Le f f ˜ Z ( Z D(Ω,x) = ∗ = ε0E(Ω,x)+g(x)Γ(Ω,x)E(Ω,x). (A42) 4 ∂E Z(J) = DA exp iSem + i d xJ · A If we then recall the usual constitutive relation for the dis- ) placement vector95, i.e., D = ε εE, we can identify the term i Z ←→ 0 + dt dt0 d3xg(x)A˙ · Γ · A˙ , (A44) Γ˜ (Ω,x) with the dielectric constant of a medium, i.e., 2 g(x) ε(Ω,x) = 1 + Γ˜ (Ω,x). (A43) ε 0 ←→ 0 where A˙ · Γ · A˙ = A˙µ (t)Γµν (t − t;,x)A˙ν (t ), and now we It should now appear clear, that Γ(t − t0,x) represents the di- have restored the current-dependent term, because now we are electric function of an effective medium the electromagnetic going to perform the path integration with respect to the vec- field is propagating through. In particular, it contains informa- tor potential A. The result above then matches the form of the tion on both the loss channels present in the medium (through effective partition function introduced in Eq. (135).