%******************************************************************************* %*********************************** Second Chapter **************************** %******************************************************************************* %Title of the Second Chapter \chapter{Quantum Optics of Quantum Gases} \ifpdf \graphicspath{{Chapter2/Figs/Raster/}{Chapter2/Figs/PDF/}{Chapter2/Figs/}} \else \graphicspath{{Chapter2/Figs/Vector/}{Chapter2/Figs/}} \fi \section{Introduction} In this chapter, we derive a general Hamiltonian that describes the coupling of atoms with far-detuned optical beams \cite{mekhov2012}. This will serve as the basis from which we explore the system in different parameter regimes, such as nondestructive measurement in free space or quantum measurement backaction in a cavity. Another interesting direction for this field of research are quantum optical lattices where the trapping potential is treated quantum mechanically. However this is beyond the scope of this work. \mynote{insert our paper citations here} We consider $N$ two-level atoms in an optical lattice with $M$ sites. For simplicity we will restrict our attention to spinless bosons, although it is straightforward to generalise to fermions, which yields its own set of interesting quantum phenomena \cite{atoms2015, mazzucchi2016, mazzucchi2016af}, and other spin particles. The theory can be also be generalised to continuous systems, but the restriction to optical lattices is convenient for a variety of reasons. Firstly, it allows us to precisely describe a many-body atomic state over a broad range of parameter values due to the inherent tunability of such lattices. Furthermore, this model is capable of describing a range of different experimental setups ranging from a small number of sites with a large filling factor (e.g.~BECs trapped in a double-well potential) to a an extended multi-site lattice with a low filling factor (e.g.~a system with one atom per site which will exhibit the Mott insulator to superfluid quantum phase transition). \mynote{extra fermion citations, Piazza? Look up Gabi's AF paper.} As we have seen in the previous section, an optical lattice can be formed with classical light beams that form standing waves. Depending on the detuning with respect to the atomic resonance, the nodes or antinodes form the lattice sites in which atoms accumulate. As shown in Fig. \ref{fig:LatticeDiagram} the trapped bosons (green) are illuminated with a coherent probe beam (red) and scatter light into a different mode (blue) which is then measured with a detector. The most straightforward measurement is to simply count the number of photons with a photodetector, but it is also possible to perform a quadrature measurement by using a homodyne detection scheme. The experiment can be performed in free space where light can scatter in any direction. The atoms can also be placed inside a cavity which has the advantage of being able to enhance light scattering in a particular direction. Furthermore, cavities allow for the formation of a fully quantum potential in contrast to the classical lattice trap. \begin{figure}[htbp!] \centering \includegraphics[width=\linewidth]{LatticeDiagram} \caption[Experimental Setup]{Atoms (green) trapped in an optical lattice are illuminated by a coherent probe beam (red), $a_0$, with a mode function $u_0(\b{r})$ which is at an angle $\theta_0$ to the normal to the lattice. The light scatters (blue) into the mode $\a_1$ in free space or into a cavity and is measured by a detector. Its mode function is given by $u_1(\b{r})$ and it is at an angle $\theta_1$ relative to the normal to the lattice. If the experiment is in free space light can scatter in any direction. A cavity on the other hand enhances scattering in one particular direction.} \label{fig:LatticeDiagram} \end{figure} For simplicity, we will be considering one-dimensional lattices most of the time. However, the model itself is derived for any number of dimensions and since none of our arguments will ever rely on dimensionality our results straightforwardly generalise to 2- and 3-D systems. This simplification allows us to present a much more intuitive picture of the physical setup where we only need to concern ourselves with a single angle for each optical mode. As shown in Fig. \ref{fig:LatticeDiagram} the angle between the normal to the lattice and the probe and detected beam are denoted by $\theta_0$ and $\theta_1$ respectively. We will consider these angles to be tunable although the same effect can be achieved by varying the wavelength of the light modes. However, it is much more intuitive to consider variable angles in our model as this lends itself to a simpler geometrical representation. \section{Derivation of the Hamiltonian} \label{sec:derivation} A general many-body Hamiltonian coupled to a quantized light field in second quantized can be separated into three parts, \begin{equation} \label{eq:FullH} \H = \H_f + \H_a + \H_{fa}. \end{equation} The term $\H_f$ represents the optical part of the Hamiltonian, \begin{equation} \label{eq:Hf} \H_f = \sum_l \hbar \omega_l \ad_l \a_l - i \hbar \sum_l \left( \eta_l^* \a_l - \eta_l \ad \right). \end{equation} The operators $\a_l$ ($\ad$) are the annihilation (creation) operators of light modes with frequencies $\omega_l$, wave vectors $\b{k}_l$, and mode functions $u_l(\b{r})$, which can be pumped by coherent fields with amplitudes $\eta_l$. The second part of the Hamiltonian, $\H_a$, is the matter-field component given by \begin{equation} \label{eq:Ha} \H_a = \int \mathrm{d}^3 \b{r} \Psi^\dagger(\b{r}) \H_{1,a} \Psi(\b{r}) + \frac{2 \pi a_s \hbar^2}{m} \int \mathrm{d}^3 \b{r} \Psi^\dagger(\b{r}) \Psi^\dagger(\b{r}) \Psi(\b{r}) \Psi(\b{r}). \end{equation} Here, $\Psi(\b{r})$ ($\Psi^\dagger(\b{r})$) are the matter-field operators that annihilate (create) an atom at position $\b{r}$, $a_s$ is the $s$-wave scattering length characterising the interatomic interaction, and $\H_{1,a}$ is the atomic part of the single-particle Hamiltonian $\H_1$. The final component of the total Hamiltonian is the interaction given by \begin{equation} \label{eq:Hfa} \H_{fa} = \int \mathrm{d}^3 \b{r} \Psi^\dagger(\b{r}) \H_{1,fa} \Psi(\b{r}), \end{equation} where $\H_{1,fa}$ is the interaction part of the single-particle Hamiltonian, $\H_1$. The single-particle Hamiltonian in the rotating-wave and dipole approximation is given by \begin{equation} \H_1 = \H_f + \H_{1,a} + \H_{1,fa}, \end{equation} \begin{equation} \H_{1,a} = \frac{\b{p}^2} {2 m_a} + \frac{\hbar \omega_a}{2} \sigma_z, \end{equation} \begin{equation} \H_{1,fa} = - i \hbar \sum_l \left[ \sigma^+ g_l \a_l u_l(\b{r}) - \sigma^- g^*_l \ad_l u^*_l(\b{r}) \right]. \end{equation} In the equations above, $\b{p}$ and $\b{r}$ are the momentum and position operators of an atom of mass $m_a$ and resonance frequency $\omega_a$. The operators $\sigma^+ = |g \rangle \langle e|$, $\sigma^- = |e \rangle \langle g|$, and $\sigma_z = |e \rangle \langle e| - |g \rangle \langle g|$ are the atomic raising, lowering and population difference operators, where $|g \rangle$ and $| e \rangle$ denote the ground and excited states of the two-level atom respectively. $g_l$ are the atom-light coupling constants for each mode. It is the inclusion of the interaction of the boson with quantized light that distinguishes our work from the typical approach to ultracold atoms where all the optical fields, including the trapping potentials, are treated classically. We will now simplify the single-particle Hamiltonian by adiabatically eliminating the upper excited level of the atom. The equations of motion for the time evolution of operator $\hat{A}$ in the Heisenberg picture are given by \begin{equation} \dot{\hat{A}} = \frac{i}{\hbar} \left[\H, \hat{A} \right]. \end{equation} Therefore, the Heisenberg equation for the lowering operator of a single particle is \begin{equation} \dot{\sigma}^- = \frac{i}{\hbar} \left[\H_1, \hat{\sigma}^- \right] = \hbar \omega_a \sigma^- + i \hbar \sum_l \sigma_z g_l \a_l u_l(\b{r}). \end{equation} We will consider nonresonant interactions between light and atoms where the detunings between the light fields and the atomic resonance, $\Delta_{la} = \omega_l - \omega_a$, are much larger than the spontaneous emission rate and Rabi frequencies $g_l \a_l$. Therefore, the atom will be predominantly found in the ground state and we can set $\sigma_z = -1$ which is also known as the linear dipole approximation as the dipoles respond linearly to the light amplitude when the excited state has negligible population. Moreover, we can adiabatically eliminate the polarization $\sigma^-$. Firstly we will re-write its equation of motion in a frame rotating at $\omega_p$, the external probe frequency, such that $\sigma^- = \tilde{\sigma}^- \exp(i \omega_p t)$, and similarly for $\tilde{\a}_l$. The resulting equation is given by \begin{equation} \dot{\tilde{\sigma}}^- = - \hbar \Delta_a \tilde{\sigma}^- - i \hbar \sum_l g_l \tilde{\a}_l u_l(\b{r}), \end{equation} where $\Delta_a = \omega_p - \omega_a$ is the atom-probe detuning. Within this rotating frame we will take $\dot{\tilde{\sigma}}^- \approx 0$ and thus obtain the following equation for the lowering operator \begin{equation} \label{eq:sigmam} \sigma^- = - \frac{i}{\Delta_a} \sum_l g_l \a_l u_l(\b{r}). \end{equation} Therefore, by inserting this expression into the Heisenberg equation for the light mode $m$ given by \begin{equation} \dot{\a}_m = - \sigma^- g_m u^*_m(\b{r}) \end{equation} we get the following equation of motion \begin{equation} \dot{\a}_m = \frac{i}{\Delta_a} \sum_l g_l g_m u_l(\b{r}) u^*_m(\b{r}) \a_l. \end{equation} An effective Hamiltonian which results in the same optical equations of motion can be written as $\H^\mathrm{eff}_1 = \H_f + \H^\mathrm{eff}_{1,a} + \H^\mathrm{eff}_{1,fa}$. The effective atomic and interaction Hamiltonians are \begin{equation} \label{eq:aeff} \H^\mathrm{eff}_{1,a} = \frac{\b{p}^2}{2 m_a} + V_\mathrm{cl}(\b{r}), \end{equation} \begin{equation} \label{eq:faeff} \H^\mathrm{eff}_{1,fa} = \frac{\hbar}{\Delta_a} \sum_{l,m} u_l^*(\b{r}) u_m(\b{r}) g_l g_m \ad_l \a_m, \end{equation} where we have explicitly extracted $V_\mathrm{cl}(\b{r}) = \hbar g_\mathrm{cl}^2 |a_\mathrm{cl} u_\mathrm{cl}(\b{r})|^2 / \Delta_{\mathrm{cl},a}$, the classical trapping potential, from the interaction terms. However, we consider the trapping beam to be sufficiently detuned from the other light modes that we can neglect any scattering between them. A later inclusion of this scattered light would not be difficult due to the linearity of the dipoles we assumed. Normally we will consider scattering of modes $a_l$ much weaker than the field forming the lattice potential $V_\mathrm{cl}(\b{r})$. Therefore, we assume that the trapping is entirely due to the classical potential and the field operators $\Psi(\b{r})$ can be expanded using localised Wannier functions of $V_\mathrm{cl}(\b{r})$. By keeping only the lowest vibrational state we get \begin{equation} \Psi(\b{r}) = \sum_i^M b_i w(\b{r} - \b{r}_i), \end{equation} where $b_i$ ($\bd_i$) is the annihilation (creation) operator of an atom at site $i$ with coordinate $\b{r}_i$, $w(\b{r})$ is the lowest band Wannier function, and $M$ is the number of lattice sites. Substituting this expression in Eq. \eqref{eq:FullH} with $\H_{1,a} = \H_{1,a}^\mathrm{eff}$ and $\H_{1,fa} = \H_{1,fa}^\mathrm{eff}$ given by Eq. \eqref{eq:aeff} and \eqref{eq:faeff} respectively yields the following generalised Bose-Hubbard Hamiltonian, $\H = \H_f + \H_a + \H_{fa}$, \begin{equation} \H = \H_f + \sum_{i,j}^M J^\mathrm{cl}_{i,j} \bd_i b_j + \sum_{i,j,k,l}^M \frac{U_{ijkl}}{2} \bd_i \bd_j b_k b_l + \frac{\hbar}{\Delta_a} \sum_{l,m} g_l g_m \ad_l \a_m \left( \sum_{i,j}^K J^{l,m}_{i,j} \bd_i b_j \right). \end{equation} The optical field part of the Hamiltonian, $\H_f$, has remained unaffected by all our approximations and is given by Eq. \eqref{eq:Hf}. The matter-field Hamiltonian is now given by the well known Bose-Hubbard Hamiltonian \begin{equation} \H_a = \sum_{i,j}^M J^\mathrm{cl}_{i,j} \bd_i b_j + \sum_{i,j,k,l}^M \frac{U_{ijkl}}{2} \bd_i \bd_j b_k b_l, \end{equation} where the first term represents atoms tunnelling between sites with a hopping rate given by \begin{equation} J^\mathrm{cl}_{i,j} = \int \mathrm{d}^3 \b{r} w (\b{r} - \b{r}_i ) \left( -\frac{\b{p}^2}{2 m_a} + V_\mathrm{cl}(\b{r}) \right) w(\b{r} - \b{r}_i), \end{equation} and $U_{ijkl}$ is the atomic interaction term given by \begin{equation} U_{ijkl} = \frac{4 \pi a_s \hbar^2}{m_a} \int \mathrm{d}^3 \b{r} w(\b{r} - \b{r}_i) w(\b{r} - \b{r}_j) w(\b{r} - \b{r}_k) w(\b{r} - \b{r}_l). \end{equation} Both integrals above depend on the overlap between Wannier functions corresponding to different lattice sites. This overlap decreases rapidly as the distance between sites increases. Therefore, in both cases we can simply neglect all terms but the one that corresponds to the most significant overlap. Thus, for $J_{i,j}^\mathrm{cl}$ we will only consider $i$ and $j$ that correspond to nearest neighbours. Furthermore, since we will only be looking at lattices that have the same separtion between all its nearest neighbours (e.g. cubic or square lattice) we can define $J_{i,j}^\mathrm{cl} = - J^\mathrm{cl}$ (negative sign, because this way $J^\mathrm{cl} > 0$). For the inter-atomic interactions this simplifies to simply considering on-site collisions where $i=j=k=l$ and we define $U_{iiii} = U$. Finally, we end up with the canonical form for the Bose-Hubbard Hamiltonian \begin{equation} \H_a = -J^\mathrm{cl} \sum_{\langle i,j \rangle}^M \bd_i b_j + \frac{U}{2} \sum_{i}^M \hat{n}_i (\hat{n}_i - 1), \end{equation} where $\langle i,j \rangle$ denotes a summation over nearest neighbours and $\hat{n}_i = \bd_i b_i$ is the atom number operator at site $i$. Finally, we have the light-matter interaction part of the Hamiltonian given by \begin{equation} \H_{fa} = \frac{\hbar}{\Delta_a} \sum_{l,m} g_l g_m \ad_l \a_m \left( \sum_{i,j}^K J^{l,m}_{i,j} \bd_i b_j \right), \end{equation} where the coupling between the matter and optical fields is determined by the coefficients \begin{equation} J^{l,m}_{i,j} = \int \mathrm{d}^3 \b{r} w(\b{r} - \b{r}_i) u_l^*(\b{r}) u_m(\b{r}) w(\b{r} - \b{r}_j). \end{equation} This contribution can be separated into two parts, one which couples directly to the on-site atomic density and one that couples to the tunnelling operators. We will define the operator \begin{equation} \label{eq:F} \hat{F}_{l,m} = \hat{D}_{l,m} + \hat{B}_{l,m}, \end{equation} where $\hat{D}_{l,m}$ is the direct coupling to atomic density \begin{equation} \label{eq:D} \hat{D}_{l,m} = \sum_{i}^K J^{l,m}_{i,i} \hat{n}_i, \end{equation} and $\hat{B}_{l,m}$ couples to the matter-field via the nearest-neighbour tunnelling operators \begin{equation} \label{eq:B} \hat{B}_{l,m} = \sum_{\langle i, j \rangle}^K J^{l,m}_{i,j} \bd_i b_j, \end{equation} where $K$ denotes a sum over the illuminated sites and we neglect couplings beyond nearest neighbours for the same reason as before when deriving the matter Hamiltonian. \mynote{make sure all group papers are cited here} These equations encapsulate the simplicity and flexibility of the measurement scheme that we are proposing. The operators given above are entirely determined by the values of the $J^{l,m}_{i,j}$ coefficients and despite its simplicity, this is sufficient to give rise to a host of interesting phenomena via measurement back-action such as the generation of multipartite entangled spatial modes in an optical lattice \cite{elliott2015, atoms2015, mekhov2009pra}, the appearance of long-range correlated tunnelling capable of entangling distant lattice sites, and in the case of fermions, the break-up and protection of strongly interacting pairs \cite{mazzucchi2016, kozlowski2016zeno}. Additionally, these coefficients are easy to manipulate experimentally by adjusting the optical geometry via the light mode functions $u_l(\b{r})$. It is important to note that we are considering a situation where the contribution of quantized light is much weaker than that of the classical trapping potential. If that was not the case, it would be necessary to determine the Wannier functions in a self-consistent way which takes into account the depth of the quantum poterntial generated by the quantized light modes. This significantly complicates the treatment, but can lead to interesting physics. Amongst other things, the atomic tunnelling and interaction coefficients will now depend on the quantum state of light. \mynote{cite Santiago's papers and Maschler/Igor EPJD} Therefore, combining these final simplifications we finally arrive at our quantum light-matter Hamiltonian \begin{equation} \label{eq:fullH} \H = \H_f -J^\mathrm{cl} \sum_{\langle i,j \rangle}^M \bd_i b_j + \frac{U}{2} \sum_{i}^M \hat{n}_i (\hat{n}_i - 1) + \frac{\hbar}{\Delta_a} \sum_{l,m} g_l g_m \ad_l \a_m \hat{F}_{l,m} - i \sum_l \kappa_l \ad_l \a_l, \end{equation} where we have phenomologically included the cavity decay rates $\kappa_l$ of the modes $a_l$. A crucial observation about the structure of this Hamiltonian is that in the interaction term, the light modes $a_l$ couple to the matter in a global way. Instead of considering individual coupling to each site, the optical field couples to the global state of the atoms within the illuminated region via the operator $\hat{F}_{l,m}$. This will have important implications for the system and is one of the leading factors responsible for many-body behaviour beyond the Bose-Hubbard Hamiltonian paradigm. Furthermore, it is also vital to note that light couples to the matter via an operator, namely $\hat{F}_{l,m}$, which makes it sensitive to the quantum state of matter. This is a key element of our treatment of the ultimate quantum regime of light-matter interaction that goes beyond previous treatments. \section{The Bose-Hubbard Hamiltonian} \subsection{The Model} The original Hubbard model was developed as a model for the motion of electrons within transition metals \cite{hubbard1963}. The key elements were the confinement of the motion to a lattice and the approximation of the Coulomb screening interaction as a simple on-site interaction. Despite these enormous simplifications the model was very succesful \cite{leggett}. The Bose-Hubbard model is an even simpler variation where instead of fermions we consider spinless bosons. It was originally devised as a toy model, but in 1998 it was shown by Jaksch \emph{et.~al.} that it can be realised with ultracold atoms in an optical lattice \cite{jaksch1998}. Shortly afterwards it was obtained in a ground-breaking experiment \cite{greiner2002}. The model has been the subject of intense research since then, because despite its simplicity it possesses highly nontrivial properties such as the superfluid to Mott insulator quantum phase transition. Furthermore, it is one of the most controllable quantum many-body systems thus providing a solid basis for new experiments and technologies. The model we have derived is essentially an extension of the well-known Bose-Hubbard model that also includes interactions with quantised light. Therefore, it should come as no surprise that if we eliminate all the quantized fields from the Hamiltonian we obtain exactly the Bose-Hubbard model \begin{equation} \H_a = -J \sum_{\langle i,j \rangle}^M \bd_i b_j + \frac{U}{2} \sum_{i}^M \hat{n}_i (\hat{n}_i - 1). \end{equation} The first term is the kinetic energy term and denotes the hopping of atoms between neighbouring sites at a hopping rate given by $J$. The second term describes the on-site repulsion which acts mainly to suppress multiple occupancies of any site, but at the same time it lies at the heart of strongly correlated phenomena in the Bose-Hubbard model. Unfortunately, despite its simplicity, this system is not solvable for any finite value $U/J$ using known techniques in finite dimensions \cite{krauth1991, kolovsky2004, calzetta2006}. However, exact solutions exist for the ground state in the two limits of $U = 0$ and $J = 0$ which already provide some insight as they correspond to the two different phases of the quantum phase transition that is present in the model. In higher dimensions (e.g.~two or three) the system is reasonably well described by a mean-field approximation which provides us with a good understanding of the full quantum phase diagram. \subsection{The Superfluid and Mott Insulator Ground States} Before trying to understand the quantum phase transition itself we will look at the ground states of the two extremal cases first and for simplicity we will consider only homogenous systems with periodic boundary conditions here. We will begin with the $U = 0$ limit where the Hamiltonian's kinetic term can be diagonalised by transforming to momentum space defined by the annihilation operator \begin{equation} b_\b{k} = \frac{1} {\sqrt{M}} \sum_m b_m e^{i \b{k} \cdot \b{r}_m}, \end{equation} where $\b{k}$ denotes the wavevector running over the first Brillouin zone. The Hamiltonian then is given by \begin{equation} \H_a = \sum_\b{k} \epsilon_\b{k} \bd_\b{k} b_\b{k}, \end{equation} where the free particle dispersion is \begin{equation} \epsilon_\b{k} = -2 J \sum_{d = 1}^D \cos(k_d a), \end{equation} and $d$ denotes the dimension out of a total of $D$ dimensions. It should now be obvious that the ground state is simply the state with all the atoms in the $\b{k} = 0$ state and thus it is given by \begin{equation} \label{eq:GSSF} | \Psi_\mathrm{SF} \rangle = \frac{ ( \bd_{\b{k} = 0} )^N } {\sqrt{N!}} | 0 \rangle = \frac{1} {\sqrt{N!}} \left( \frac{1} {\sqrt{M}} \sum_m^M \bd_m \right)^N | 0 \rangle, \end{equation} where $| 0 \rangle$ denotes the vacuum state. This state is called a superfluid (SF) due to its viscous-free flow properties \cite{leggett1999}. The macroscopic occupancy of the single-particle ground state is the signature and defining property of a Bose-Einstein condensate. Note that the zero momentum states consist of a superposition of an atom at all the sites of the optical lattice highlighting the fact that kinetic energy is minimised by delocalising the particles. This in turn implies that a superfluid state will have infinite range correlations. It is also important to note that in the thermodynamic limit the superfluid state is gapless, i.e.~it takes zero energy to excite the system. At the other end of the spectrum, i.e.~$J = 0$, the kinetic term vanishes and the system is dominated by the on-site interactions. This has the effect of decoupling the Hamiltonian into a sum of identical single-site Hamiltonians which are already diagonal in the atom number basis. Therefore, the ground state for $N = gM$, where $g$ is an integer, is given by \begin{equation} \label{eq:GSMI} | \Psi_\mathrm{MI} \rangle = \prod_m^M \frac{1} {\sqrt{g!}} (\bd_m)^g | 0 \rangle. \end{equation} This state has precisely the same number of bosons, $g$, at each site and is commonly known as the Mott insulator (MI). Unlike in the superfluid state, the atoms are all localised to a specific lattice site and thus not only are there no long-range correlations, there are actually no two-site correlations in this state at all. Additionally, this state is gapped, i.e.~it costs a finite amount of energy to excite the system into its lowest excited state even in the thermodynamic limit. In fact, the existance of the gap is the defining property that distinguishes the Mott insulator from the superfluid in the Bose-Hubbard model. Note that we have only considered a commensurate case (number of atoms an integer multiple of the number of sites). If we were to insert another atom on top of the Mott insulator, the energy of the system would not depend on the lattice site in which it is inserted. This implies that as soon as $J$ becomes finite, the ground state would prefer this atom be delocalised and effectively becomes gapless which means the system is a superfluid and it will not exhibit a quantum phase transition. \subsection{Mean-Field Theory} Now that we have a basic understanding of the two limiting cases we can now consider the model in between these two extremes. We have already mentioned that an exact solution is not known, but fotunately a very good mean-field approximation exists \cite{fisher1989}. In this approach the interaction term is treated exactly, but the kinetic energy term is decoupled as \begin{equation} \bd_m b_n = \langle \bd_m \rangle b_n + \bd_m \langle b_n \rangle - \langle \bd_m \rangle \langle b_n \rangle = \Phi^* b_n + \Phi \bd_m - | \Phi |^2, \end{equation} where the expectation values are for the $T = 0$ ground state. Note that we have assumed that $\Phi = \langle b_m \rangle $ is non-zero and homogenous. $\Phi$ describes the influence of hopping between neighbouring sites and is the mean -field order parameter. This decoupling effect means that we can write the Hamiltonian as $\hat{H}_a = \sum_m^M \hat{h}_a$, where \begin{equation} \hat{h}_a = -z J ( \Phi \bd \Phi^* b) + z J | \Phi |^2 + \frac{U}{2} \n (\n - 1) - \mu \n, \end{equation} and $z$ is the coordination number, i.e. the number of nearest neighbours of a single site. We have also introduced the chemical potential $\mu$ since we now consider the grand canonical ensemble as the Hamiltonian no longer conserves the total atom number. The ground state of the $| \Psi_0 \rangle$ of the overall system will be a site-wise product of the individual ground states of $\hat{h}_a$. These can be found very easily using standard diagonalisation techniques. $\Phi$ is then self-consistently determined by minimising the energy of the ground state with respect to $\Phi$. The main advantage of the mean-field treatment is that it lets us study the quantum phase transition between the sueprfluid and Mott insulator phases discussed in the previous sections. The phase boundaries can be obtained from second-order perturbation theory as \begin{equation} \left( \frac{\mu} {U} \right)_\pm = \frac{1}{2} \left[ 2g - 1 - \left( \frac{zJ}{U} \right) \pm \sqrt{ 1 - 2 (2g + 1) \left( \frac{zJ}{U} \right) + \left( \frac{zJ}{U} \right)^2} \right]. \end{equation} This yields a phase diagram with multiple Mott insulating lobes as seen in Fig. \ref{fig:BHPhase} where each lobe represents a Mott insualtor with a different number of atoms per site. The tip of the lobe which corresponds to the quantum phase transition along a line of fixed density is obtained by equating the two branches of the above equation to get \begin{equation} \left( \frac{J} {U} \right) = \frac{1} {z (2g + 1 + \sqrt{4g^2 + 4g} )}. \end{equation} It is important to note that the fact that this formalism predicts a phase transition is in contrast to other mean-field theories such as the Bogoliubov approximation \cite{PitaevskiiStringari}. This highlights the fact that the interactions, which are treated exactly here, are the dominant driver leading to this phase transition and other strongly correlated effects in the Bose-Hubbard model. \begin{figure} \centering \includegraphics[width=0.8\linewidth]{BHPhase} \caption[Mean-Field Bose-Hubbard Phase Diagram]{Mean-field phase diagram of the Bose-Hubbard model in 1D, i.e.~ $z = 2$, from Ref. \cite{StephenThesis}. The shaded regions are the Mott insulator lobes and each lobe corresponds to a different on-site filling labeled by $n$. The rest of the space corresponds to the superfluid phase. The dashed lines are the phase boundaries obtained from first-order perturbation theory. The solid lines are are lines of constant density in the superfluid phase. \label{fig:BHPhase}} \end{figure} \subsection{The Bose-Hubbard Model in One Dimension} The mean-field theory in the previous section is very useful tool for studying the quantum phase transition in the Bose-Hubbard model. However, it is effectively an infinite-dimensional theory and in practice it only works in two dimensions or more. The phase transition in 1D is poorly described, because it actually belongs to a different universality class. This is clearly seen from the one dimensional phase diagram shown in Fig. \ref{fig:1DPhase}. \begin{figure} \centering \includegraphics[width=0.8\linewidth]{1DPhase} \caption[Exact 1D Bose-Hubbard Phase Diagram]{Exact phase diagram of the Bose-Hubbard model in 1D, i.e.~ $z = 2$, from Ref. \cite{StephenThesis}. The shaded region corresponds to the $n = 1$ Mott insulator lobe. The sharp tip distinguishes it from the mean-field phase diagram. The red dotted line denotes the reentrance phase transition which occurs for a fixed $\mu$.The critical point for the fixed density $n = 1$ transition is denoted by an 'x'. \label{fig:1DPhase}} \end{figure} Some general conclusions can be obtained by looking at Haldane's prescription for Luttinger liquids \cite{haldane1981, giamarchi}. Without a periodic potential the low-energy physics of the system is described by the Hamiltonian \begin{equation} \hat{H}_a = \frac{1}{2 \pi} \int \mathrm{d} x \left\{ v K [ \hat{\Pi}(x) ]^2 + \frac{v} {K} [\partial_x \hat{\Phi}(x) ]^2 \right\}, \end{equation} where we have expressed the bosonic field operators in terms of a density operator $\hat{\rho}(x)$ and a phase operator $\hat{\Phi}(x)$ as $\hat{\Psi}(x) = \sqrt{\hat{\rho}(x)} e^{i \hat{\hat{\Phi}(x)}}$ and $\hat{\Pi}(x)$ is the density fluctuation operator. Provided the parameters $v$ and $K$ can be correctly determined this Hamiltonian gives the correct description of the gapless superfluid phase of the Bose-Hubbard model. Most importantly it gives an expression for the spatial correlation functions such as \begin{equation} \langle \bd_m b_n \rangle = A \left( \frac{\alpha} {|m - n|} \right)^{K/2}, \end{equation} where $A$ is some amplitude and $\alpha$ is a necessary cutoff to regularise the theory at short distances. Unlike the superfluid ground state in Eq. \eqref{eq:GSSF} this state does not have infinite range correlations. They decay according to a power law. However, for non-interacting systems $K = 0$ and long-range order is re-established as before though it is important to note that in higher dimensions this long-range order persists in the whole superfluid phase even with interactions present. In order to describe the phase transition and the Mott insulating phase it is necessary to introduce a periodic lattice potential. It can be shown that this system exhibits at $T = 0$ a Berezinskii-Kosterlitz-Thouless phase transition as the parameter $K$ is varied with a critical point at $K = \frac{1}{2}$ where $K < \frac{1}{2}$ is a superfluid. Above $K = \frac{1}{2}$ the value of $K$ jumps discontinuously to $K \rightarrow \infty$ producing the Mott insulator phase. Unlike the gapless superfluid phase the spatial correlations decay exponentially as \begin{equation} \langle \bd_m b_n \rangle = B e^{ - |m - n|/\xi}, \end{equation} where $B$ is some constant and the correlation length is given by $\xi = v / \Delta$ where $\Delta$ is the energy gap. Using advanced numerical methods such as density matrix renormalisation group (DMRG) calculations it is possible to identify the critical point by fitting the power-law decay correlations in order to obtain $K$. The resulting phase transition is shown in Fig. \ref{fig:1DPhase} and the critical point was shown to be $(U/zJ) = 1.68$. Note that unusually the phase diagram exhibits a reentrance phase transition for a fixed $\mu$. \section{Scattered light behaviour} \label{sec:a} Having derived the full quantum light-matter Hamiltonian we will now look at the behaviour of the scattered light. We begin by looking at the equations of motion in the Heisenberg picture \begin{equation} \dot{\a_l} = \frac{i}{\hbar} [\hat{H}, \a_l] = -i \left(\omega_l + U_{l,l}\hat{F}_{1,1} \right) \a_l - i \sum_{m \ne l} U_{l,m} \hat{F}_{l,m} \a_m - \kappa_l \a_l + \eta_l, \end{equation} where we have defined $U_{l,m} = g_l g_m / \Delta_a$. By considering the steady state solution in the frame rotating with the probe frequency $\omega_p$ we can show that the stationary solution is given by \begin{equation} \a_l = \frac{\sum_{m \ne l} U_{l,m} \a_m \hat{F}_{l,m} + i \eta_l} {(\Delta_{lp} - U_{l,l} \hat{F}_{l,l}) + i \kappa_l}, \end{equation} where $\Delta_{lp} = \omega_p - \omega_l$ is the detuning between the probe beam and the light mode $a_l$. This equation gives us a direct relationship between the light operators and the atomic operators. In the stationary limit, the quantum state of the light field adiabatically follows the quantum state of matter. The above equation is quite general as it includes an arbitrary number of light modes which can be pumped directly into the cavity or produced via scattering from other modes. To simplify the equation slightly we will neglect the cavity resonancy shift, $U_{l,l} \hat{F}_{l,l}$ which is possible provided the cavity decay rate and/or probe detuning are large enough. We will also only consider probing with an external coherent beam, $a_0$ with mode function $u_0(\b{r})$, and thus we neglect any cavity pumping $\eta_l$. We also limit ourselves to only a single scattered mode, $a_1$ with a mode function $u_1(\b{r})$. This leads to a simple linear relationship between the light mode and the atomic operator $\hat{F}_{1,0}$ \begin{equation} \label{eq:a} \a_1 = \frac{U_{1,0} a_0} {\Delta_{p} + i \kappa} \hat{F} = C \hat{F}, \end{equation} where we have defined $C = U_{1,0} a_0 / (\Delta_{p} + i \kappa)$ which is essentially the Rayleigh scattering coefficient into the cavity. Furthermore, since there is no longer any ambiguity in the indices $l$ and $m$, we have dropped indices on $\Delta_{1p} \equiv \Delta_p$, $\kappa_1 \equiv \kappa$, and $\hat{F}_{1,0} \equiv \hat{F}$. We also do the same for the operators $\hat{D}_{1,0} \equiv \hat{D}$, $\hat{B}_{1,0} \equiv \hat{B}$, and the coefficients $J^{1,0}_{i,j} \equiv J_{i,j}$. The operator $\a_1$ itself is not an observable. However, it is possible to combine the outgoing light field with a stronger local oscillator beam in order to measure the light quadrature \begin{equation} \hat{X}_\phi = \frac{1}{2} \left( \a_1 e^{-i \phi} + \ad_1 e^{i \phi} \right), \end{equation} which in turn can expressed via the quadrature of $\hat{F}$, $\hat{X}^F_\beta$, as \begin{equation} \hat{X}_\phi = |C| \hat{X}_\beta^F = \frac{|C|}{2} \left( \hat{F} e^{-i \beta} + \hat{F}^\dagger e^{i \beta} \right), \end{equation} where $\beta = \phi - \phi_C$, $C = |C| \exp(i \phi_C)$, and $\phi$ is the local oscillator phase. Whilst the light amplitude and the quadratures are only linear in atomic operators, we can easily have access to higher moments via related quantities such as the photon number $\ad_1 \a_1 = |C|^2 \hat{F}^\dagger \hat{F}$ or the quadrature variance \begin{equation} \label{eq:Xvar} ( \Delta X_\phi )^2 = \langle \hat{X}_\phi^2 \rangle - \langle \hat{X}_\phi \rangle^2 = \frac{1}{4} + |C|^2 (\Delta X^F_\beta)^2, \end{equation} which reflect atomic correlations and fluctuations. Finally, even though we only consider a single scattered mode, this model can be applied to free space by simply varying the direction of the scattered light mode if multiple scattering events can be neglected. This is likely to be the case since the interactions will be dominated by photons scattering from the much larger coherent probe. \section{Density and Phase Observables} \label{sec:B} Light scatters due to its interactions with the dipole moment of the atoms which for off-resonant light, like the type that we consider, results in an effective coupling with atomic density, not the matter-wave amplitude. Therefore, it is challenging to couple light to the phase of the matter-field as is typical in quantum optics for optical fields. Most of the exisiting work on measurement couples directly to atomic density operators \cite{mekhov2012, LP2009, rogers2014, ashida2015, ashida2015a}. However, it has been shown that one can couple to the interference term between two condensates (e.g.~a BEC in a double-well) by using interference measurements \cite{cirac1996, castin1997, ruostekoski1997, ruostekoski1998, rist2012}. Such measurements establish a relative phase between the condensates even though the two components have initially well-defined atom numbers which is phase's conjugate variable. In a lattice geometry, one would ideally measure between two sites similarly to single-site addressing \cite{greiner2009, bloch2011}, which would measure a single term $\langle \bd_i b_{i+1}+b_i \bd_{i+1}\rangle$. This could be achieved, for example, by superposing a deeper optical lattice shifted by $d/2$ with respect to the original one, catching and measuring the atoms in the new lattice sites. However, a single-shot success rate of atom detection will be small and as single-site addressing is challenging, we proceed with our global scattering scheme. In our model light couples to the operator $\hat{F}$ which consists of a density component, $\hat{D} = \sum_i J_{i,i} \hat{n}_i$, and a phase component, $\hat{B} = \sum_{\langle i, j \rangle} J_{i,j} \bd_i b_j$. In general, the density component dominates, $\hat{D} \gg \hat{B}$, and thus $\hat{F} \approx \hat{D}$. However, it is possible to engineer an optical geometry in which $\hat{D} = 0$ leaving $\hat{B}$ as the dominant term in $\hat{F}$. This approach is fundamentally different from the aforementioned double-well proposals as it directly couples to the interference terms caused by atoms tunnelling rather than combining light scattered from different sources. Furthermore, it is not limited to a double-wellsetup and naturally extends to a lattice structure which is a key advantage. Such a counter-intuitive configuration may affect works on quantum gases trapped in quantum potentials \cite{mekhov2012, mekhov2008, larson2008, chen2009, habibian2013, ivanov2014, caballero2015} and quantum measurement-induced preparation of many-body atomic states \cite{mazzucchi2016, mekhov2009prl, pedersen2014, elliott2015}. \mynote{add citiations above if necessary} For clarity we will consider a 1D lattice as shown in Fig. \ref{fig:LatticeDiagram} with lattice spacing $d$ along the $x$-axis direction, but the results can be applied and generalised to higher dimensions. Central to engineering the $\hat{F}$ operator are the coefficients $J_{i,j}$ given by \begin{equation} \label{eq:Jcoeff} J_{i,j} = \int \mathrm{d} x \,\,\, w(x - x_i) u_1^*(x) u_0(x) w(x - x_j). \end{equation} The operators $\hat{B}$ and $\hat{D}$ depend on the values of $J_{i,i+1}$ and $J_{i,i}$ respectively. These coefficients are determined by the convolution of the coupling strength between the probe and scattered light modes, $u_1^*(x)u_0(x)$, with the relevant Wannier function overlap shown in Fig. \ref{fig:WannierProducts}a. For the $\hat{B}$ operator we calculate the convolution with the nearest neighbour overlap, $W_1(x) \equiv w(x - d/2) w(x + d/2)$, and for the $\hat{D}$ operator we calculate the convolution with the square of the Wannier function at a single site, $W_0(x) \equiv w^2(x)$. Therefore, in order to enhance the $\hat{B}$ term we need to maximise the overlap between the light modes and the nearest neighbour Wannier overlap, $W_1(x)$. This can be achieved by concentrating the light between the sites rather than at the positions of the atoms. \mynote{Potentially expand details of the derivation of these equations} In order to calculate the $J_{i,j}$ coefficients we perform numerical calculations using realistic Wannier functions \cite{walters2013}. However, it is possible to gain some analytic insight into the behaviour of these values by looking at the Fourier transforms of the Wannier function overlaps, $\mathcal{F}[W_{0,1}](k)$, shown in Fig. \ref{fig:WannierProducts}b. This is because for plane and standing wave light modes the product $u_1^*(x) u_0(x)$ can be in general decomposed into a sum of oscillating exponentials of the form $e^{i k x}$ making the integral in Eq. \eqref{eq:Jcoeff} a sum of Fourier transforms of $W_{0,1}(x)$. We consider both the detected and probe beam to be standing waves which gives the following expressions for the $\hat{D}$ and $\hat{B}$ operators \begin{align} \label{eq:FTs} \hat{D} = & \frac{1}{2}[\mathcal{F}[W_0](k_-)\sum_m\hat{n}_m\cos(k_- x_m +\varphi_-) \nonumber\\ & + \mathcal{F}[W_0](k_+)\sum_m\hat{n}_m\cos(k_+ x_m +\varphi_+)], \nonumber\\ \hat{B} = & \frac{1}{2}[\mathcal{F}[W_1](k_-)\sum_m\hat{B}_m\cos(k_- x_m +\frac{k_-d}{2}+\varphi_-) \nonumber\\ & +\mathcal{F}[W_1](k_+)\sum_m\hat{B}_m\cos(k_+ x_m +\frac{k_+d}{2}+\varphi_+)], \end{align} where $k_\pm = k_{0x} \pm k_{1x}$, $k_{(0,1)x} = k_{0,1} \sin(\theta_{0,1}$), $\hat{B}_m=b^\dag_mb_{m+1}+b_mb^\dag_{m+1}$, and $\varphi_\pm=\varphi_0 \pm \varphi_1$. The key result is that the $\hat{B}$ operator is phase shifted by $k_\pm d/2$ with respect to the $\hat{D}$ operator since it depends on the amplitude of light in between the lattice sites and not at the positions of the atoms allowing to decouple them at specific angles. \begin{figure}[hbtp!] \centering \includegraphics[width=0.8\linewidth]{BDiagram} \caption[Maximising Light-Matter Coupling between Lattice Sites]{Light field arrangements which maximise coupling, $u_1^*u_0$, between lattice sites. The thin black line indicates the trapping potential (not to scale). (a) Arrangement for the uniform pattern $J_{i,i+1} = J_1$. (b) Arrangement for spatially varying pattern $J_{i,i+1}=(-1)^m J_2$; here $u_0=1$ so it is not shown and $u_1$ is real thus $u_1^*u_0=u_1$. \label{fig:BDiagram}} \end{figure} \begin{figure}[hbtp!] \centering \includegraphics[width=\linewidth]{WF_S} \caption[Wannier Function Products]{The Wannier function products: (a) $W_0(x)$ (solid line, right axis), $W_1(x)$ (dashed line, left axis) and their (b) Fourier transforms $\mathcal{F}[W_{0,1}]$. The Density $J_{i,i}$ and matter-interference $J_{i,i+1}$ coefficients in diffraction maximum (c) and minimum (d) as are shown as functions of standing wave shifts $\varphi$ or, if one were to measure the quadrature variance $(\Delta X^F_\beta)^2$, the local oscillator phase $\beta$. The black points indicate the positions, where light measures matter interference $\hat{B} \ne 0$, and the density-term is suppressed, $\hat{D} = 0$. The trapping potential depth is approximately 5 recoil energies.} \label{fig:WannierProducts} \end{figure} The simplest case is to find a diffraction maximum where $J_{i,i+1} = J^B_\mathrm{max}$, where $J^B_\mathrm{max}$ is a constant. This results in a diffraction maximum where each bond (inter-site term) scatters light in phase and the operator is given by \begin{equation} \hat{B} = J^B_\mathrm{max} \sum_m^K \hat{B}_m . \end{equation} This can be achieved by crossing the light modes such that $\theta_0 = -\theta_1$ and $k_{0x} = k_{1x} = \pi/d$ and choosing the light mode phases such that $\varphi_+ = 0$. Fig. \ref{fig:BDiagram}a shows the resulting light mode functions and their product along the lattice and Fig. \ref{fig:WannierProducts}c shows the value of the $J_{i,j}$ coefficients under these circumstances. In order to make the $\hat{B}$ contribution to light scattering dominant we need to set $\hat{D} = 0$ which from Eq. \eqref{eq:FTs} we see is possible if \begin{equation} \xi \equiv \varphi_0 = -\varphi_1 = \frac{1}{2}\arccos\left[\frac{-\mathcal{F}[W_0](2\pi/d)}{\mathcal{F}[W_0](0)}\right]. \end{equation} Under these conditions, the coefficient $J^B_\mathrm{max}$ is simply given by $J^B_\mathrm{max} = \mathcal{F}[W_1](2 \pi / d)$. This arrangement of light modes maximizes the interference signal, $\hat{B}$, by suppressing the density signal, $\hat{D}$, via interference compensating for the spreading of the Wannier functions. Another possibility is to obtain an alternating pattern similar corresponding to a diffraction minimum where each bond scatters light in anti-phase with its neighbours giving \begin{equation} \hat{B} = J^B_\mathrm{min} \sum_m^K (-1)^m \hat{B}_m, \end{equation} where $J^B_\mathrm{min}$ is a constant. We consider an arrangement where the beams are arranged such that $k_{0x} = 0$ and $k_{1x} = \pi/d$ which gives the following expressions for the density and interference terms \begin{align} \label{eq:DMin} \hat{D} = & \mathcal{F}[W_0]\left(\frac{\pi}{d}\right) \sum_m (-1)^m \hat{n}_m \cos(\varphi_0) \cos(\varphi_1) \nonumber \\ \hat{B} = & -\mathcal{F}[W_1]\left(\frac{\pi}{d}\right) \sum_m (-1)^m \hat{B}_m \cos(\varphi_0) \sin(\varphi_1). \end{align} For $\varphi_0 = 0$ the corresponding $J_{i,j}$ coefficients are given by $J_{i,i+1} = (-1)^i J^B_\mathrm{min}$, where $J^B_\mathrm{min} = -\mathcal{F}[W_1](\pi / d)$, and are shown in Fig. \ref{fig:WannierProducts}d. The light mode coupling along the lattice is shown in Fig. \ref{fig:BDiagram}b. It is clear that for $\varphi_1 = \pm \pi/2$, $\hat{D} = 0$, which is intuitive as this places the lattice sites at the nodes of the mode $u_1(x)$. This is a diffraction minimum as the light amplitude is also zero, $\langle \hat{B} \rangle = 0$, because contributions from alternating inter-site regions interfere destructively. However, the intensity $\langle \ad_1 \a \rangle = |C|^2 \langle \hat{B}^2 \rangle$ is proportional to the variance of $\hat{B}$ and is non-zero. Alternatively, one can use the arrangement for a diffraction minimum described above, but use travelling instead of standing waves for the probe and detected beams and measure the light quadrature variance. In this case $\hat{X}^F_\beta = \hat{D} \cos(\beta) + \hat{B} \sin(\beta)$ and by varying the local oscillator phase, one can choose which conjugate operator to measure. \section{Electric Field Stength} \label{sec:Efield} The Electric field operator at position $\b{r}$ and at time $t$ is usually written in terms of its positive and negative components: \begin{equation} \b{\hat{E}}(\b{r},t) = \b{\hat{E}}^{(+)}(\b{r},t) + \b{\hat{E}}^{(-)}(\b{r},t), \end{equation} where \begin{subequations} \begin{equation} \b{\hat{E}}^{(+)}(\b{r},t) = \sum_{\b{k}} \hat{\epsilon}_{\b{k}} \mathcal{E}_{\b{k}} \a_{\b{k}} e^{-i \omega_{\b{k}} t + i \b{k} \cdot \b{r}}, \end{equation} \begin{equation} \b{\hat{E}}^{(-)}(\b{r},t) = \sum_{\b{k}} \hat{\epsilon}_{\b{k}} \mathcal{E}_{\b{k}} \a^\dagger_{\b{k}} e^{i \omega_{\b{k}} t - i \b{k} \cdot \b{r}}, \end{equation} \end{subequations} $\hat{\epsilon}_{\b{k}}$ is a unit polarization vector, $\b{k}$ is the wave vector, $\mathcal{E}_{\b{k}} = \sqrt{\hbar \omega_{\b{k}} / 2 \epsilon_0 V}$, $\epsilon_0$ is the free space permittivity, $V$ is the quantization volume and $\a_\b{k}$ and $\a_\b{k}^\dagger$ are the annihilation and creation operators respectively of a photon in mode $\b{k}$, and $\omega_\b{k}$ is the angular frequency of mode $\b{k}$. In Ref. \cite{Scully} it was shown that the operator $\b{\hat{E}}^{(+)}(\b{r},t)$ due to light scattering from a single atom located at $\b{r}^\prime$ at the observation point $\b{r}$ is given by \begin{equation} \label{eq:Ep} \b{\hat{E}}^{(+)}(\b{r},\b{r}^\prime,t) = \frac{\omega_a^2 d_A \sin \eta}{4 \pi \epsilon_0 c^2 |\b{r} - \b{r}^\prime|} \hat{\epsilon} \sigma^- \left( \b{r}^\prime, t - \frac{|\b{r} - \b{r}^\prime|}{c} \right), \end{equation} with a similar expression for $\b{E}^{(-)}(\b{r},\b{r}^\prime,t)$. Eq. \eqref{eq:Ep} is valid only in the far field, $\eta$ is the angle the dipole makes with $\b{r} - \b{r}^\prime$, $\hat{\epsilon}$ is the polarization vector which is perpendicular to $\b{r} - \b{r}^\prime$ and lies in the plane defined by $\b{r} - \b{r}^\prime$ and the dipole, $\omega_a$ is the atomic transition frequency, and $d_A$ is the dipole matrix element between the two levels, and $c$ is the speed of light in vacuum. We have already derived an expression for the atomic lowering operator, $\sigma^-$, in Eq. \eqref{eq:sigmam} and it is given by \begin{equation} \sigma^-(\b{r}, t) = - \frac{i} {\Delta_a} \sum_\b{k} g_\b{k} a_\b{k}(t) u_\b{k}(\b{r}). \end{equation} We will want to substitute this expression into Eq. \eqref{eq:Ep}, but first we note that in the setup we consider the coherent probe will be much stronger than the scattered modes. This in turn implies that the expression for $\sigma^-$ will be dominated by this external probe. Therefore, we drop the sum and consider only a single wave vector, $\b{k}_0$, of the dominant contribution from the external beam corresponding to the mode $\a_0$. We now use the definititon of the atom-light coupling constant \cite{Scully} to make the substitution $g_0 a_0 = -i \Omega_0 e^{-i \omega_0 t} / 2$, where $\Omega_0$ is the Rabi frequency for the probe-atom system. We now evaluate the polarisation operator at the retarded time \begin{equation} \sigma^- \left(\b{r}^\prime, t - \frac{|\b{r} - \b{r}^\prime|}{c} \right) = -\frac{\Omega_0}{2 \Delta_a} u_0 (\b{r}^\prime) \exp \left[-i \omega_0 \left( t - \frac{|\b{r} - \b{r}^\prime|}{c} \right)\right]. \end{equation} We are considering observation in the far field regime so $|\b{r} - \b{r}^\prime| \approx r - \hat{\b{r}} \cdot \b{r}^\prime$. We also consider the incoming, $\b{k}_0$, and the outgoing, $\mathbf{k}_1$, waves to be of the same wavelength, $|\b{k}_0| = |\b{k}_1| = k = \omega_0 / c$, hence $(\omega_0 / c) |\b{r} - \b{r}^\prime| \approx k (r - \hat{\b{r}} \cdot \b{r}^\prime) = \b{k}_1 \cdot (\b{r} - \b{r}^\prime )$ and \begin{equation} \sigma^- \left(\b{r}^\prime, t - \frac{|\b{r} - \b{r}^\prime|}{c} \right) \approx -\frac{\Omega_0}{2 \Delta_a} u_0 (\b{r}^\prime) \exp \left[i \b{k}_1 \cdot (\b{r} - \b{r}^\prime ) - i \omega_0 t \right]. \end{equation} The many-body version of the electric field operator is given by \begin{equation} \b{\hat{E}}^{(+)}_N(\b{r},t) = \int \mathrm{d}^3 \b{r}^\prime \Psi^\dagger (\b{r}^\prime) \b{\hat{E}}^{(+)}(\b{r},\b{r}^\prime,t) \Psi(\b{r}^\prime), \end{equation} where $\Psi(\b{r})$ is the atomic matter-field operator. As before, we expand this field operator using localized Wannier functions corresponding to the lattice potential and keeping only the lowest vibrational state at each site: $\Psi(\b{r}) = \sum_i b_i w(\b{r} - \b{r}_i)$, where $b_i$ is the annihilation operator of an atom at site $i$ with the coordinate $\b{r}_i$. Thus, the relevant many-body operator is \begin{subequations} \begin{equation} \b{\hat{E}}^{(+)}_N(\b{r},t) = \hat{\epsilon} C_E \sum_{i,j}^K \bd_i b_j \int \mathrm{d}^3 \b{r}^\prime w^* (\b{r}^\prime - \b{r}_i) \frac{u_0 (\b{r}^\prime)}{|\b{r} - \b{r}^\prime|} e^ {i \b{k}_1 \cdot (\b{r} - \b{r}^\prime ) - i \omega_0 t } w(\b{r}^\prime - \b{r}_j), \end{equation} \begin{equation} C_E = -\frac{\omega_a^2 \Omega_0 d \sin \eta}{8 \pi \epsilon_0 c^2 \Delta_a} \end{equation} \end{subequations} where $K$ is the number of illuminated sites. We now assume that the lattice is deep and that the light scattering occurs on time scales much shorter than atom tunneling. Therefore, we ignore all terms for which $i \ne j$ and the remaining integrals become $\int \mathrm{d}^3 \b{r}_0 w^*(\b{r}_0 - \b{r}_i) f (\b{r}) w(\b{r}_0 - \b{r}_i) = f(\b{r}_i)$. The final form of the many body operator is then \begin{equation} \b{E}^{(+)}_N(\b{r},t) = \hat{\epsilon} C_E \sum_{j = 1}^K \hat{n}_j \frac{u_0 (\b{r}_j)}{|\b{r} - \b{r}_j|} e^ {i \b{k}_1 \cdot (\b{r} - \b{r}_j ) - i \omega_0 t }, \end{equation} where $\hat{n}_i = b^\dagger_ib_i$ is the atom number operator at site $i$. We will now consider the incoming probe to be a plane wave, $u_0(\b{r}) = e^{i \b{k}_0 \cdot \b{r}}$, and we approximate $|\b{r} - \b{r}_j| \approx r$ in the denominator. Thus, \begin{equation} \b{E}^{(+)}_N(\b{r},t) = \hat{\epsilon} C_E \frac{e^ {ikr - i\omega_0t}}{r} \sum_{j = 1}^K \hat{n}_j e^ {i (\b{k}_0 - \b{k}_1) \cdot \b{r}_j}. \end{equation} We note that this is exactly the same form of the optical field as derived from a generalised Bose-Hubbard model which considers a fully quantum light-matter interaction Hamiltonian \cite{mekhov2007prl, mekhov2007pra, mekhov2012}. We can even express the result in terms of the $\hat{D} = \sum_j^K u_1^*(\b{r}_j) u_0(\b{r}_j) \hat{n}_j$ formalism used in those works \begin{equation} \b{E}^{(+)}_N(\b{r},t) = \hat{\epsilon} C_E \frac{e^ {ikr - i\omega_0t}}{r} \hat{D}. \end{equation} The average light intensity at point $\b{r}$ at time $t$ is given by the formula $I = c \epsilon_0 |E|^2/2$ and yields \begin{equation} \langle I (\b{r}, t) \rangle = c \epsilon_0 \langle E^{(-)} (\b{r}, t) E^{(+)} (\b{r}, t) \rangle = \frac{c \epsilon_0}{r^2} C_E^2 \langle \hat{D}^* \hat{D} \rangle. \end{equation} The scattering is dominated by a uniform background for which $\langle \hat{D}^* \hat{D} \rangle \approx N_K$ for a superfluid \cite{mekhov2012}, where $N_K$ is the mean number of atoms in the illuminated volume. Thus, the approximate number of photons scattered per second can be obtained by integrating over a sphere \begin{equation} n_{\Phi} = \frac{4 \pi c \epsilon_0}{3 \hbar \omega_a} \left(\frac{\omega_a^2 \Omega_0 d}{8 \pi \epsilon_0 c^2 \Delta_a}\right)^2 N_K. \end{equation} In the Weisskopf-Wigner approximation it can be shown \cite{Scully} that for a two level system the decay constant, $\Gamma$, is \begin{equation} \Gamma = \frac{\omega_a^3 d^2}{3 \pi \epsilon_0 \hbar c^3}. \end{equation} Therefore, we can now express the quantity $n_{\Phi}$ as \begin{equation} n_{\Phi} = \frac{1}{8} \left(\frac{\Omega_0}{\Delta_a}\right)^2 \frac{\Gamma}{2} N_K. \end{equation} Estimates of the scattering rate using real experimental parameters are given in Table \ref{tab:photons}. Rubidium atom data has been taken from Ref. \cite{steck}. Miyake \emph{et al.} experimental parameters are from Ref. \cite{miyake2011}. The $5^2S_{1/2}$, $F=2 \rightarrow 5^2P_{3/2}$, $F^\prime = 3$ transition of $^{87}$Rb is considered. For this transition the Rabi frequency is actually larger than the detuning and and effects of saturation should be taken into account in a more complete analysis. However, it is included for discussion. The detuning was specified as ``a few linewidths'' and note that it is much smaller than $\Omega_0$. The Rabi fequency is $\Omega_0 = (d_\mathrm{eff}/\hbar)\sqrt{2 I / c \epsilon_0}$ which is obtained from definition of Rabi frequency, $\Omega = \mathbf{d} \cdot \mathbf{E} / \hbar$, assuming the electric field is parallel to the dipole, and the relation $I = c \epsilon_0 |E|^2 /2$. We used $d_\mathrm{eff} = 1.73 \times 10^{-29}$ Cm \cite{steck}. Weitenberg \emph{et al.} experimental parameters are based on Ref. \cite{weitenberg2011, weitenbergThesis}. The $5^2S_{1/2} \rightarrow 5^2P_{3/2}$ transition of $^{87}$Rb is used. Ref. \cite{weitenberg2011} gives the free space detuning to be $\Delta_\mathrm{free} = - 2 \pi \cdot 45$ MHz, but Ref. \cite{weitenbergThesis} clarifies that the relevant detuning is $\Delta = \Delta_\mathrm{free} + \Delta_\mathrm{lat}$, where $\Delta_\mathrm{lat} = - 2 \pi \cdot 40$ MHz. Ref. \cite{weitenbergThesis} uses a saturation parameter $s_\mathrm{tot}$ to quantify the intensity of the beams which is related to the Rabi frequency, $s_\mathrm{tot} = 2 \Omega^2 / \Gamma^2$ \cite{steck,foot}. We can extract $s_\mathrm{tot}$ for the experiment from the total scattering rate by $\Gamma_\mathrm{sc} = (\Gamma/2) (s_\mathrm{tot}) / (1+s_\mathrm{tot}+(2 \Delta / \Gamma)^2)$. A scattering rate of 60 kHz per atom \cite{weitenberg2011} gives $s_\mathrm{tot} = 2.5$. \begin{table} \centering \begin{tabular}{l c c} \toprule Value & Miyake \emph{et al.} & Weitenberg \emph{et al.} \\ \midrule $\omega_a$ & \multicolumn{2}{ c }{$2 \pi \cdot 384$ THz}\\ $\Gamma$ & \multicolumn{2}{ c }{$2 \pi \cdot 6.07$ MHz} \\ $\Delta_a$ & $2\pi \cdot 30$ MHz & $2 \pi \cdot 85$ MHz \\ $I$ & $4250$ Wm$^{-2}$ & N/A \\ $\Omega_0$ & 293$\times 10^6$ s$^{-1}$ & 42.5$\times 10^6$ s$^{-1}$ \\ $N_K$ & 10$^5$ & 147 \\ \midrule $n_{\Phi}$ & $6 \times 10^{11}$ s$^{-1}$ & $2 \times 10^6$ s$^{-1}$ \\ \bottomrule \end{tabular} \caption[Photon Scattering Rates]{Experimental parameters used in estimating the photon scattering rates.} \label{tab:photons} \end{table}