diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzjskk" "b/data_all_eng_slimpj/shuffled/split2/finalzzjskk" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzjskk" @@ -0,0 +1,5 @@ +{"text":"\\section{Introduction}\n\nNGC 5548 is a bright, low-redshift (z $=$ 0.0172) Seyfert 1 galaxy that has \nreceived a considerable amount of attention over the past decade. In particular, \nNGC 5548 has been the subject of a number of intensive spectroscopic monitoring \ncampaigns in the optical and UV (Korista et al. 1995). \nThese efforts have yielded important results on the nature of the continuum \nsource and the size, geometry, and kinematics of the broad-line region (BLR), \nfor which the responsivity peaks at about 20 light days from the nucleus\n(Peterson et al. 1994). Little attention has been paid to the narrow-line region \n(NLR) in NGC 5548, however, although it could lead to a better understanding of \nthe circumnuclear environment at much greater distances from the nucleus of this\notherwise well-studied active galaxy.\n\nIn general, studies of the NLR in active galaxies are important for \nunderstanding the nature of the NLR clouds and the interaction of the central \ncontinuum source with the surrounding galaxy on large scales. Emission-line \nstudies and detailed photoionization modeling are particularly useful \nfor determining the range of physical conditions and reddening amongst the NLRs \nin active galaxies. A comparison of the NLR properties in Seyfert \n1 and Seyfert 2 galaxies should be helpful in testing unified theories, which \npostulate that the two types are the same object viewed from \ndifferent perspectives, such that the continuum source and BLR are ``hidden'' in \nSeyfert 2 galaxies (Miller \\& Goodrich 1990; Antonucci 1993). If this basic \nhypothesis is correct, then the intrinsic properties of the NLRs in Seyfert 1 \nand Seyfert 2 galaxies should {\\it not} show large systematic differences.\n\nThe narrow lines in Seyfert 1 galaxies are more difficult to measure \nthan in Seyfert 2 galaxies, due to blending with the broad lines. However, given \nspectra with sufficient signal-to-noise ratio and spectral resolution, these \ncomponents can be isolated and measured reasonably well\n(Crenshaw \\& Peterson 1986). Measurements of the narrow lines in the optical are \ngiven by Cohen (1983) for a large number of Seyfert 1 galaxies. In the UV, these \nmeasurements were difficult with the low spectral resolution of the \n{\\it International Ultraviolet Explorer} ({\\it IUE}), but are possible with \ninstruments on \nthe {\\it Hubble Space Telescope} ({\\it HST}), such as the Faint Object \nSpectrograph (FOS).\n\nThe FOS UV spectra of NGC 5548 presented in a previous paper (Crenshaw, Boggess, \n\\& Wu 1993; hereafter Paper I) provide a good starting point for detailed \nstudies of the NLR in a Seyfert 1 galaxy. These observations happened to occur\nat a time when the broad emission lines (and continuum fluxes) were at an \nhistoric low in the UV, and the contrast between the broad and narrow \ncomponents is thereby enhanced. Paper I gives the UV spectrum and measurements \nof the broad and narrow lines in NGC 5548. Relative to the other narrow lines, \nC~IV $\\lambda$1549 is much stronger in NGC 5548 than in Seyfert 2 galaxies, \nindicating a higher ionization parameter and\/or harder continuum in the NLR of \nNGC 5548. Narrow Mg II $\\lambda$2800 emission is very weak or absent in NGC \n5548, and Paper I presents two possible explanations: 1) the NLR clouds lack the \npresence of a partially-ionized zone (i.e., they are optically thin to ionizing \nradiation), and\/or 2) dust grains are present in the NLR clouds, and the Mg II \nflux is weak due to depletion and\/or destruction from multiple scatterings and \neventual absorption of the photons by dust (Kraemer \\& Harrington 1986; Ferland \n1992).\n\nWe now have the opportunity to investigate the preliminary results from Paper I \nin more detail, by including ground-based optical spectra and photoionization \nmodels. From the ground-based monitoring campaigns, we have selected \nspectra that cover the full optical range (3000 -- 10,000 \\AA) and were\nobserved around the same time period as the UV data, when the broad-line fluxes \nwere very low. The combination of optical and UV lines provides a wide range of \nemission-line diagnostics, as well as an opportunity to deredden the lines using \nthe He~II recombination lines. We can then use multicomponent \nphotoionization models to match the dereddened line \nratios and probe the physical conditions in the NLR of NGC 5548.\n\n\\section{Observations and Data Analysis}\n\n\\subsection{UV and Optical Spectra}\n\nWe obtained the FOS UV spectra of NGC 5548 through a 1\\arcsecpoint0 circular \naperture on 1992 July 5 UT. Paper I gives the details of the observations and \nmeasurements, along with the UV spectrum and emission-line fluxes (with \nassociated errors). The observations were made prior to the installation of \nCOSTAR on {\\it HST}, so near-simultaneous {\\it IUE} spectra were used to adjust \nthe absolute flux levels of the FOS spectra. As noted in Paper I, \nthe scale factors needed to bring the FOS continuum fluxes up to the {\\it IUE} \nlevels are around 1.4 -- 1.5, which are somewhat higher than the \nvalues of 1.1 -- 1.3 for our other Seyfert observations. We concluded that the \nSeyfert nucleus may not have been accurately centered in the aperture. Koratkar \net al. (1996) suggest that another possible explanation for the discrepancy in \nabsolute fluxes is nonlinearity in the {\\it IUE} detectors. However, we have \nseen no evidence for this possibility in other observations at these flux \nlevels, so we have no reason to distrust the {\\it IUE} fluxes. In addition, \nreprocessed versions of these spectra that we obtained from the {\\it HST} and \n{\\it IUE} archives have not changed the original fluxes by more than 10\\%, so we \ncontinue to use the values from Paper I. Some of the emission lines in Paper I \nhave only a single number quoted for the flux (as opposed to separate values for \nthe broad and narrow components); a single value represents the narrow-line \ncontribution, since the broad component is either not present or too weak to be \ndetected in these cases.\n\nWe selected two optical spectra obtained during a four-year monitoring campaign \non NGC 5548 (Peterson et al. 1994), from a time interval of $\\sim$30 days when \nthe H$\\beta$ and continuum light curves were at their lowest levels to date.\nThe spectra were chosen on the basis of their large wavelength coverage\n(3000 - 10000 \\AA), high signal-to-noise ratio ($\\geq$50 per resolution \nelement in the continuum at 5200 \\AA), and acceptable resolution ($\\sim$8 \n\\AA). The spectra were obtained through a 4\\arcsecpoint0 x \n10\\arcsecpoint0 aperture with the 3.0-m Shane telescope $+$ Kast spectrograph on \n1992 April 21 and 1992 May 23 UT. Additional details on the observations are \ngiven by Peterson et al. (1994). The absolute flux levels were \nadjusted by scaling the optical spectra so that the [O~III] $\\lambda$5007 flux \nis 5.58 x 10$^{-13}$ ergs s$^{-1}$ cm$^{-2}$, a value determined from \nobservations through large apertures on spectrophotometric nights (Peterson et \nal. 1991). The scale factors we used are 1.36 for the 1992 April 24 spectrum and \n1.01 for the 1992 May 23 spectrum.\n\nPlots of the optical spectra are shown in Figure 1 (the UV spectrum is shown in \nPaper I). The contrast between the broad and narrow components of the permitted \nemission lines is most clearly seen in H$\\beta$. The 1992 April 21 spectrum was \nobtained at an historic low level, with a continuum flux of \nF$_{\\lambda}$(5100~\\AA) $=$ 5.5 x 10$^{-15}$ ergs s$^{-1}$ \ncm$^{-2}$ \\AA$^{-1}$ and total H$\\beta$ flux of F(H$\\beta$) $=$ 3.2 x 10$^{-13}$ \nergs s$^{-1}$ cm$^{-2}$. The flux levels are a little higher for the 1992 May 23 \nspectrum with F$_{\\lambda}$(5100 \\AA) $=$ 6.0 x 10$^{-15}$ ergs s$^{-1}$ \ncm$^{-2}$ \\AA$^{-1}$ and F(H$\\beta$) $=$ 3.7 x 10$^{-13}$ ergs s$^{-1}$ \ncm$^{-2}$. At the time of the FOS \nobservations on 1992 July 5, the continuum and H$\\beta$ fluxes were close to the \nsame levels as those from the second optical spectrum, according to the light \ncurves of Peterson et al. (1994).\n\nAlthough the optical aperture is much larger than the one used for the UV \nobservations, we have substantial evidence that it does not contain much \nadditional NLR flux. Unfortunately, there are no {\\it HST} narrow-band images in \n[O~III] or other strong lines that could be used to directly determine the \ndistribution of narrow-line emission close to the nucleus. However, there is \nsignificant evidence that the apparent size of the NLR is very small in NGC \n5548. Peterson et al. (1995) find from a ground-based image that the [O~III] \nemission is pointlike, given a point-spread function that is characterized by a \nwidth of 2\\arcsecpoint0 (FWHM). In addition, Wilson \\& Ulvestad (1982) show that \nin an aperture that is 4\\arcsecpoint2 in diameter, the [O~III] $\\lambda$5007 \nfluxes at positions offset from the nucleus by 4\\arcsecpoint5 -- 6$''$\nare about 100 times weaker than the nuclear flux.\nMore importantly, in Paper I we found that the strongest UV lines in the {\\it \nIUE} 20$''$ x 10$''$ aperture have fluxes that are only $\\sim$ 20\\% higher than \nthose in the FOS 1\\arcsecpoint0 aperture. Thus, the observed UV to optical line \nratios that we quote are at most 20\\% too low, which has little effect on our \ncomparisons with the model results.\n\nIn order to measure the flux of each narrow optical line, we used\na local baseline determined by linear interpolation between adjacent \ncontinuum regions or broad profile wings (in the case of profiles \nconsisting of broad and narrow components). For severely blended lines like \nH$\\alpha$ and [N~II] $\\lambda\\lambda$6548, 6584, we used the [O~III] \n$\\lambda$5007 profile as a template to deblend the lines (see Crenshaw \\& \nPeterson 1986). The adopted flux for each narrow component is the average of the \nvalues from each of the two spectra.\n\nWe determined the reddening of the narrow emission lines from the He~II \n$\\lambda$1640\/$\\lambda$4686 ratio and the Galactic reddening curve of Savage \\& \nMathis (1979). For the temperatures and densities typical of the NLR, the He II \nlines are due to recombination, and this particular ratio only varies from 6.3 \nto 7.6 (Seaton 1978); we adopt an intrinsic value of 7.2, consistent with our \nmodel values (Section 3). The observed He~II $\\lambda$1640\/$\\lambda$4686 ratio \nis 5.5 $\\pm$ 1.6, which yields a reddening of E$_{B-V}$ $=$ 0.07 mag \n$^{+0.09}_{-0.06}$. The portion of the reddening that is due to our own Galaxy \nis E$_{B-V}$ $=$ 0.03 mag, determined from a neutral hydrogen column density of \nN$_{HI}$ $=$ 1.6 x 10$^{20}$ cm$^{-2}$ (Murphy et al. 1996) and the relationship \nE$_{B-V}$ $=$N$_{HI}$\/5.2 x 10$^{21}$ cm$^{-2}$ (Shull \\& Van Steenburg 1985).\nWe note that the intrinsic reddening of the narrow emission lines in this \nSeyfert 1 galaxy, E$_{B-V}$ $\\approx$ 0.04 mag, is much smaller than typical \nvalues of 0.2 -- 0.4 mag obtained for Seyfert 2 galaxies (MacAlpine 1988; \nFerland \\& Osterbrock 1986; Kraemer et al. 1994). \n\nWe determined errors in the dereddened ratios from the sum in quadrature of the \nerrors from three sources: photon noise, different reasonable continuum \nplacements, and reddening. Errors in the optical ratios are dominated by \ncontinuum placement, whereas errors in the UV to optical ratios are due to both \ncontinuum placement and uncertainties in the reddening correction. Errors in the \nweak lines in both regions also have a significant contribution from photon \nnoise. As we discussed earlier in this section, there are some possible sources \nof systematic error in the UV to optical line ratios, on which we placed upper \nlimits of $\\sim$20\\%.\n\nTable 1 gives the observed and dereddened narrow-line ratios relative to \nH$\\beta$, and errors in the dereddened ratios. Cohen (1983) gives the next \nmost comprehensive list of optical line ratios; in general, Cohen's observed \nratios agree with ours to within the errors. A number of investigators have \nindependently determined the narrow H$\\beta$\/[O~III] $\\lambda$5007 ratio in NGC \n5548 (Cohen 1983; Crenshaw \\& Peterson 1986; Peterson 1987; Wamsteker et al. \n1990; Rosenblatt et al. 1992; Wanders \\& Peterson 1996): these values range from \n0.10 to 0.15, compared to our value of 0.12 $\\pm$ 0.01.\n\n\\subsection{The Ionizing Continuum}\n\nEstimates of the ionizing continuum are needed as input values for the \nphotoionization models of the NLR. We choose the continuum data points given by \nKrolik et al. (1991), since they represent the historic mean levels for this \nobject. As always, the greatest uncertainty is the shape of the extreme \nultraviolet (EUV) continuum. Figure 2 \ngives the UV continuum point closest to the EUV region, at 1340 \\AA, and the \nX-ray continuum points from Krolik et al. (cf. Turner \n\\& Pounds 1989; Clavel et al. 1991) in terms of luminosity (ergs \ns$^{-1}$ Hz$^{-1}$), which we have adjusted for a Hubble constant of H$_{0}$ $=$ \n75 km s$^{-1}$ Mpc$^{-1}$. The dotted line in Figure 1 gives Krolik et al.'s \ncontinuum fit in the EUV, which is a power law determined from the UV data along \nwith an exponential cutoff designed to meet the first X-ray point. We prefer a \nfit with two power laws (in the form L$_{\\nu}$ $=$ K$\\nu^{\\alpha}$), given the \nevidence for an upturn in the spectrum at energies smaller than 1 -- 2 keV \n(i.e., a soft X-ray excess). A fit to these data yields $\\alpha$ $=$ \n$-$1.40$\\pm$0.03 in the EUV and soft X-rays, and $\\alpha$ $=$ $-$0.40$\\pm$0.03 \nin the hard X-rays; the break point is at $\\nu$ $=$ 10$^{17.1}$ Hz$^{-1}$ (1.3 \nkeV).\n\nNGC 5548 was monitored by the {\\it Extreme Ultraviolet Explorer} ({\\it EUVE}) \nover a two month period during 1993 March -- May \n(Marshall et al. 1997). During this time, the EUV flux varied by a factor of \nfour from peak to minimum, and the average flux (corrected for Galactic \nneutral hydrogen absorption) was 135 $\\mu$Jy at $\\sim$76 \\AA.\nThese observations provide an important constraint on the EUV ionizing \ncontinuum, but were not used directly in our continuum fit for two reasons. \nFirst, the neutral hydrogen absorption due to our Galaxy is well known (see the \nprevious section), but there could be additional absorption along the \nline of sight. Second, the {\\it EUVE} flux, averaged over two months, may not \nbe representative of the average flux over many years. Given these caveats, we \nplot the {\\it EUVE} continuum point in Figure 2 for comparison with our adopted \ncontinuum; the error bar was determined from an estimate of $\\pm$10\\% \nuncertainty in \nGalactic N$_{HI}$ (see Murphy et al. 1996). The {\\it EUVE} point is slightly \nhigher than \nthe continuum fits in Figure 2, but appears to be consistent with our and Krolik \net al.'s \nadoption of a relatively steep continuum. If we use Krolik et al.'s continuum or \na continuum formed by joining the UV, {\\it EUVE}, and X-ray points with line \nsegments \n(in log space), the total EUV flux increases by factors of only 1.19 and 1.30, \nrespectively, and the flux at the frequency of the {\\it EUVE} observation \nincreases by \nfactors of 1.40 and 1.84, respectively. The effects of adopting these other \ncontinua are small, and will be discussed later in the paper.\n\n\\section{Photoionization Models}\n\nIn modeling the narrow-line emission of NGC 5548, we have adhered to our basic \nphilosophy of keeping the number of free parameters to a minimum, by using the \navailable observational constraints and the simplest assumptions possible.\nThe parameters are varied until the best agreement is obtained with the \nobserved line ratios, and additional input parameters are only included if \nthey are needed to provide a reasonable match to the majority of the \nlines. Discrepancies between the model predictions and specific lines are then \ninvestigated to provide further insight into the physical conditions. In some \ncases we generate variations on the standard model using additional parameters \n(such as dust) or nonstandard values of the initial parameters (such as nonsolar \nabundances) to illustrate our ideas for resolving the discrepancies.\n\n\\subsection{Methodology}\n\nThe basic modeling methodology that we employ is described in \nKraemer et al. (1994) and the details of the photoionization code are given in \nKraemer (1985). To review the major points, we assume plane parallel \ngeometry, which is reasonable if the gas is ionized by radiation from a central \nsource at a distance that is large compared to the extent of the cloud. The gas \nis assumed to be atomic (i.e., there is no molecular component). For radiation \nbounded models, we stop the integration into the slab when the electron \ntemperature falls below 5000 K and there is no longer any significant line \nemission. The emission line photon escape is through the ionized face of the \nslab. Details of the treatment of dust in the models are described in Kraemer \n(1985). Since the work of Kraemer et al. (1994), we have added iron to the \nelements modeled in the code. The atomic data that we used can be found in \nPradhan \\& Peng (1994), and references contained therein, as well as through \nFerland's ``Cloudy and Associates'' World Wide Web site \n(http:\/\/www.pa.uky.edu\/$\\sim$gary\/cloudy).\nThe final output of these models is an emission line spectrum. The line\nstrengths are tabulated relative to H$\\beta$. In addition, the model\ngives the volume emissivity of H$\\beta$, from which we can determine\nthe mass of gas required to produce the observed line emission, the\nefficiency of production of H$\\beta$ photons, and an estimate of the covering \nfactor.\n\nIn order to keep the input parameters to a minimum, we kept two of them fixed \nfor our standard model: the shape of the ionizing continuum and the abundances. \nWe used the simplest possible ionizing continuum consistent with the \nobservations, as described in Section 2.2. In addition, we have assumed \nsolar elemental abundances for the standard model as follows (see Lambert \\& \nLuck 1978): He = 0.1, C = 3.4 x 10$^{-4}$, O = 6.8 x 10$^{-4}$, N = 1.2 x \n10$^{-4}$, Ne = 1.1 x 10$^{-4}$, S = 1.5 x 10$^{-4}$, Si = 3.1 x 10$^{-5}$, Mg = \n3.3 x 10$^{-5}$, Fe = 4 x 10$^{-5}$ (relative to hydrogen by number).\n\nOur photoionization models are parameterized in terms of the density of\natomic hydrogen (N$_{H}$) and the dimensionless ionization parameter at the \nilluminated face of the cloud:\n\n\\begin{equation}\nU = \\int^{\\infty}_{\\nu_0} ~\\frac{L_\\nu}{h\\nu}~d\\nu ~\/~ (4\\pi~D^2~N_{H}~c),\n\\end{equation}\n\n\\noindent where {\\it L$_{\\nu}$} is the frequency-dependent luminosity of \nthe ionizing continuum, {\\it D} is the distance between the cloud and the \nionizing source, and h$\\nu_{0}$ = 13.6 eV. \n\nWe show that we must add two enhancements to our standard model to obtain an \nacceptable match to the observations. First, we need two components of gas, \ncharacterized by different ionization parameters and densities. Second, we show \nthat the inner component must be optically thin (i.e., radiation bounded) at the \nLyman edge (13.6 eV). We are then able to vary the ionization parameter and \ndensity of each component to match the observations. Of course, the resulting \nstandard model is an oversimplification, since it is likely that the NLR clouds \nare characterized by a number of different ionization parameters, densities, and \noptical depths. We have effectively averaged the initial conditions for each of \nthese components to fit the largest selection of line ratios. Given this \nsimplification, the two-component model gives a surprisingly good fit \nto the observations.\n\nThe narrow-line spectrum of NGC 5548 is dominated by high ionization lines,\nsuch as C~IV $\\lambda$1549, N V $\\lambda$1240, and [Ne V] $\\lambda\\lambda$\n3346, 3426, as well as the coronal lines of [Fe~VII] and [Fe~X]. This indicates \nthat there is a high ionization component relatively near the central source, \nbut presumably outside the BLR, since these lines are much narrower\n($\\leq$~500~km~s$^{-1}$ FWHM) than the broad lines ($\\sim$ 5000 km s$^{-1}$ \nFWHM).\nAlso present in the spectrum are relatively strong lines of [N II] \n$\\lambda\\lambda$6584, 6548 and [O II] $\\lambda$3727. These are typical of the \nnarrow emission line regions of many Seyfert 2 galaxies (Koski 1978; Shuder\n\\& Osterbrock 1981), and presumably arise in a component of relatively low \nionization and low density.\n\nThe choice of density for these models is based on the relative strengths\nof certain forbidden emission lines in the spectrum of NGC 5548. \nAn emission line will not be an important coolant in gas with sufficiently \nlarge electron density that collisional de-excitation of the line will dominate\nover radiative transition (see Osterbrock 1989). For example, the observed \n[Ne~V]~$\\lambda$3426\/H$\\beta$\nratio indicates that the density of the component in which that emission\noriginates must be less than $\\sim$5~x~10$^{7}$ cm$^{-3}$ (DeRobertis \\& \nOsterbrock 1984). Given the hardness of the ionizing continuum, one would expect \nhigh temperatures in this highly ionized gas.\nThe lower limit on the density of this component can be estimated from the\nrelative strength of the [O~III] $\\lambda$5007 emission. The ratio of [O III] \n$\\lambda$5007\/ [Ne V] $\\lambda$3426\nis less than 5, which indicates either low density, highly ionized gas, as\nis often seen in the extended NLR of Seyfert 2 galaxies (Storchi-\nBergmann et al. 1996), or density greater than 10$^{6}$ cm$^{-3}$ and \nsome collisional supression of the $\\lambda$5007 line. The ratio of [O~III] \n$\\lambda$4363\/$\\lambda$5007 is very high ($\\sim$0.09). In the low density \nlimit, this ratio is\n$\\sim$7 x 10$^{-3}$ at a temperature of 10$^{4}$~K (Osterbrock 1989). \nAt this low density, it is unlikely that \ntemperatures consistent with photoionization equilibrium can increase this\nratio to the observed value, and probable that the large \n$\\lambda$4363\/$\\lambda$5007\nratio is due to the fact that some of this emission arises in gas of high\ndensity (i.e., $>$ 10$^{6}$ cm$^{-3}$). \nAt this density and level of ionization, the \n[N~II]~ $\\lambda\\lambda$6548, 6584 emission from this component will be \nnegligible, so there must be a lower ionization and lower density component \npresent. In order for the [N~II] lines to be among the principal coolants from \nthis component, its density must be less than 1 x 10$^{5}$ cm$^{-3}$. Other \nstudies (cf. Filippenko \\& Halpern 1984; Filippenko 1985; Kraemer et al. 1994) \nhave shown that a range of densities in the NLR of Seyfert galaxies is likely, \nso it is not surprising that this condition exists in NGC 5548.\n\nTo summarize, the range in density, along with the presence of emission lines \nfrom a wide range of ionization states, indicates that more than one model \ncomponent is needed to fit the NLR spectrum of NGC 5548. \nThe existence of strong high ionization lines such as C IV \n$\\lambda$1549, N V $\\lambda$1240, and [Ne V] $\\lambda\\lambda$3346, 3426, and our \nevidence for high densities in the region that they are produced, requires a \ncomponent of gas relatively close to the central source. The weakness of Mg II \n$\\lambda$2800 and [O I] $\\lambda\\lambda$6300, 6364 indicate that these gas \nclouds lack a significant partially ionized zone, and therefore must be\n{\\it optically thin} to the ionizing radiation (or ``matter bounded''). \nWe further investigate the conditions in these two components below.\n\n\\subsection {Model Results and Comparison to Observations}\n\nOur approach in modeling NGC 5548 was to fit the high ionization\ncomponent first and then add components as needed to fit the lower\nionization lines (in the end, only one additional component was needed). Given \nthe constraints and assumptions described in the\nprevious section, we arrived at values of N$_{H}$ $=$ 1 x 10$^{7}$ cm$^{-3}$ and\nU $=$ 10$^{-1.5}$ for the high ionization component. Substantially lower \ndensities would result in \n[O~III]~$\\lambda$5007 being too strong, and higher densities would quench the \n[Ne V] emission. A higher ionization parameter is possible but, given our EUV \ncontinuum, would not increase the relative strengths of any of the high \nionization lines other than [Fe X] $\\lambda$6374, at the expense of putting this \ncomponent at distances much closer than $\\sim$1 pc from the continuum source \n(see Section 4). Models were run to varying optical depth at the Lyman limit \n$\\tau_{0}$, with the constraint that the Mg II and [O I] lines could not become \ntoo strong. After comparing the results of models run with $\\tau_{0}$ $=$ 1.5 to \n10, we found that $\\tau_{0}$ $=$ 2.5 gave the best fit. The emission line \nspectrum from this model, INNER, is given in Table 2.\n\nA second component, OUTER, was needed to fit the lower ionization lines. \nWe found that U $=$ 10$^{-2.5}$ and N$_{H}$ $=$ 2 x 10$^{4}$ cm$^{-3}$ gave a \ngood simultaneous fit to the [O~III]~$\\lambda$5007\/[O~II]~$\\lambda$3727 and \n[N~II] $\\lambda$6584\/H$\\beta$ ratios. Unlike our models for Mrk 3 and I~Zw~92 \n(Kraemer \\& Harrington 1986; Kraemer et al. 1994), there was no need to add a \nthird component for NGC 5548, since there is no obvious contribution from a \ncomponent of very low density ($<$ 10$^{3}$ cm$^{-3}$) low \nionization gas, such as very strong [O II] $\\lambda$3727 and [N~I] $\\lambda$5200 \nlines. The resulting emission line spectrum from OUTER is also included in \nTable 2. (We will discuss two variations on INNER and OUTER in the next \nsection.)\n\nIn order to fit the observed (and dereddened) narrow-line spectrum of NGC 5548, \nwe combined the output spectrum of the two standard components INNER and OUTER. \nIn previous studies, we attempted to weigh the contributions from each component \nto fit specific emission line ratios. For the model of NGC 5548, we simply took \nan equal contribution from INNER and OUTER. The relative simplicity\nof the narrow-line spectrum and lack of a strong contribution from very \nlow-ionization gas makes such a simple fit possible. The combined spectrum\nis given, along with the dereddened observed spectrum for comparison, in\nTable 3 (horizontal lines in the table indicate that the models do not predict \nthe strengths of these emission lines). \n\nComparison of the model predictions to the dereddened observed spectrum in Table \n3 shows agreement, to within the errors, for most of the lines. In particular, \nthese include C~IV~$\\lambda$1549, He II \n$\\lambda$1640, [O~II] $\\lambda$3727, [Ne III] $\\lambda$3869, [N II] \n$\\lambda$6584, and the Balmer decrement.\nIn radiation bounded gas,\nthe ratio of the He II lines to H$\\beta$ is strongly dependent on the shape of \nthe ionizing continuum, because neutral hydrogen is the \ndominant absorber of ionizing radiation between 13.6 eV and 54.4 eV, while above \n54.4 eV, singly ionized helium dominates.\nIf there is a component of matter bounded gas, He II\/H$\\beta$ is less easily \npredicted. The accuracy of our fit to this ratio indicates that the relative \ncontributions of the matter and radiation bounded components are approximately \ncorrect, given the observational constraints on the ionizing continuum.\nThe fact that we have a reasonable fit \nfor lines that span a wide range of ionization and critical densities supports \nour values for density and ionization parameter. Most of the \ndiscrepancies between the observations and models are in the lowest and highest \nionization lines, which we will address below. \n\n\\subsection{Discrepancies and Possible Explanations}\n\nFirst, we address differences between the predicted and observed ratios for the \nlow ionization lines. The Mg II $\\lambda$2800 and [O I] $\\lambda\\lambda$6300, \n6364 lines are still predicted to be too strong by our standard model, by \nfactors $\\geq$ 6 and 2.5, respectively. Nearly all this emission is coming from \nOUTER. Two factors determine the strength of the [O I] lines: the hardness of \nthe ionizing continuum and the physical depth of the emission line clouds. Since \nit appears that we have a good fit for the ionizing continuum, the weakness in \nthe observed [O~I] lines gives a limit on the depth of the clouds. Truncating \nthe integration of OUTER at $\\tau_{0}$ $\\approx$ 1000 would give a better fit to \nthe [O I] without affecting the other important line ratios. This results in a \ncloud depth of $\\sim$2.5 x 10$^{15}$ cm.\nThe overprediction of the Mg II $\\lambda$2800 line strength presents a somewhat\ndifferent problem. In order to reduce the contribution of this line from INNER, \nwe assumed a matter bounded model for this component. To provide a better match \nto the observation of little or no Mg II emission, we can reduce the model \ncontribution by modifying OUTER. The Mg$^{+}$ emissivity is greatest near the \nH$^{+}$\/H$^{0}$ transition zone in OUTER, so a \nsimple truncation at much lower optical depths is not feasible, as it would have\na much greater effect on the other line ratios. \n\nAn obvious explanation for the weak observed Mg II is depletion of the magnesium \ninto dust grains, along with suppression of the resonance photons by multiple \nscatterings and eventual absorption by dust. This was suggested in Paper I (cf. \nKraemer \\& Harrington 1986; Ferland 1992). For comparison with our standard \nmodel, we generated a version of OUTER that includes dust, assuming \na dust to gas ratio that is 30\\% of that found in the Galactic interstellar \nmedium, with equal amounts of graphite and silicate grains and accompanying \ndepletions. These assumptions were made to avoid biasing our results by simply \nhaving all of the Mg depleted into dust grains. We assumed relative element \ndepletions as calculated by Seab \\& \nShull (1983), and the grain size distributions determined by Mathis et al. \n(1977) and Draine \\& Lee (1984); details of the treatment of dust in the code \nare given by Kraemer (1985). The results of the model are given in Table 2, and \nnot only show a significant drop in the relative strength of Mg II \n$\\lambda$2800, but also a drop in the Ly$\\alpha$ strength, as is expected due to \nthe preferential dust absorption of multiply scattered UV resonance lines. The \nlower Ly$\\alpha$\/H$\\beta$ ratio is a better fit to the observations. A \nsubstantially larger dust-to-gas ratio than we assumed would result in a \nLy$\\alpha$\/H$\\beta$ ratio that is lower than observed. Therefore, \nit is likely that there is some dust mixed in with the low-ionization gas, \nalthough with a lower dust to gas ratio than found in the ISM, and that \ndepletion coupled with the resonance line suppression explain the weak Mg II. \n\nSecond, we address discrepancies in the high ionization lines.\nSpecifically, the lines of N~V, [Ne~V], [Fe~VII], and [Fe~X] are too weak by \nfactors of 2 to 4 compared to the observations.\nAs we mentioned earlier, the density is well constrained, so increasing the \nionization parameter brings the gas well within 1 pc, into the realm of \nthe BLR. However, these lines are relatively narrow (FWHM $\\leq$ 500 km \ns$^{-1}$, see Moore et al. 1996) and they are not likely to arise very close to \nthe BLR. The strengths of the high ionization lines can also be enhanced \nrelative to H$\\beta$ by truncating the integration of INNER at a lower optical \ndepth. However, this has the problem of enhancing the He II emission relative to \nH$\\beta$ in the model. More likely solutions to the problem of underpredicting \nthe high ionization lines include 1) shock ionization, 2) a large ``blue \nbump'' in the EUV continuum, or 3) supersolar abundances.\n\nPredicting the strengths of the coronal lines has always been a problem with \nsimple photoionization models, as Viegas-Aldrovandi \\& Contini (1989) discuss in \nsome detail for the Fe lines. They suggest that there may be shocked gas mixed \nin with the photoionized clouds and that these high ionization lines may arise \nthere. Although this is certainly a possible factor, there may be other \nplausible explanations which avoid adding another level of complexity.\n\nAn obvious way in which the coronal lines might be enhanced is if there\nwere a component of ionizing radiation that contributed significantly\nat energies between 100 and 500 eV. Although there has been some\nspeculation about the presence of a ``blue bump'' in the EUV, recent work\nby Zheng et al. (1997) on low-redshift quasars shows that the {\\it near} EUV \ncontinuum is likely to be much steeper than previously \nsupposed. In NGC 5548, the {\\it EUVE} continuum point in Figure 2 is further \nconfirmation that a large blue bump is not present in the spectrum \nof NGC 5548. Another possibility may be diffuse radiation from the intercloud \nmedium. Tran (1995) has shown that there may be a contribution to the continuum \nradiation in some Seyfert 2 galaxies from thermal emission from the intercloud\nmedium responsible for the scattering of the hidden BLR emission into\nthe observer's line of sight. If a similar medium with temperatures\n$\\approx$ 5~x~10$^{5}$~K exists in the NLR of NGC 5548, it is possible that \nfree-free radiation and line emission that arise within it may contribute to the \nionization of the inner narrow-line gas. Although this component would be weak \ncompared to the continuum radiation emitted by the central source, it could have\na significant local contribution to the ionization balance of clouds existing\nwithin this inner region. However, recent observations suggest that the extended \nUV continuum seen in some Seyferts may be due to starbursts (Heckman et al. \n1997), which would not contribute to the high ionization lines.\n\nIn studies of medium redshift QSOs, Ferland et al. (1996) found\nevidence of supersolar abundances. There is no direct evidence of\nelemental enhancements in Seyfert galaxies, but it is certainly not\nimplausible, particularly near the nucleus, where the most intense activity\noccurs. As Oliva (1996) points out, the coronal line emission will be enhanced \nproportionally to the abundance of the atomic species.\nAs a comparison, we ran a version of INNER with a heavy element\nabundance that is twice solar, and the results are shown in Table 3. In \nparticular,\nthe relative strengths of the [Fe VII] and [Ne V] lines have increased,\nwhile many lines, such as C IV $\\lambda$1549 and C III] $\\lambda$1909, show \nlittle change. It is possible, then, that the observed strength of \nsome of these lines is in part due to enhanced abundances. Note that we have not \nattempted to adjust the increase in abundances to fit assumptions about the\ntype of star formation that might be expected.\n\n\\section{Discussion}\n\nFrom our standard model, we can estimate several global properties of the NLR in \nNGC 5548, including the covering factor and physical size, in addition to more \nlocal properties, including optical depth and presence of dust.\nWe determine the covering factor of the NLR gas from the observed and model \nvalues of the ``conversion efficiency'', $\\eta$ , which is the ratio of H$\\beta$ \nphotons to ionizing photons. The covering factor is given by\nC~$=$~$\\eta$(observed)\/$\\eta$(model). A value of C $>$ 1 would indicate that the \nionizing radiation is anisotropic, which we would not expect to be the case \nfor a Seyfert 1 galaxy, since the central source is seen directly in such \nobjects. Assuming Ho= 75 km s$^{-1}$ Mpc$^{-1}$, the observed H$\\beta$ \nflux corresponds to a luminosity of 3.8 x 10$^{40}$ ergs s$^{-1}$, or 9.3 x \n10$^{51}$ H$\\beta$ photons s$^{-1}$. From the continuum observations described \nin section 2.2, we calculate a total luminosity of ionizing photons of 1.09 x \n10$^{54}$ s$^{-1}$. This yields an observed $\\eta$ = 0.009. Our fit to the \nobserved emission line spectrum assumed that each component in our model \ncontributed 50\\% of the H$\\beta$ emission. The resulting values of $\\eta$ were \n0.06 for INNER and 0.11 for OUTER, and the covering factors are 0.07 and 0.04, \nrespectively, so C(NLR) = 0.11. The value is small compared to those found for \nSeyfert 2 galaxies, which are often $>$ 1 (Kinney et al. 1991; Kraemer et al. \n1994), and there is no evidence for anisotropic radiation. Note that this \nestimate of covering factor does not include the BLR clouds, which contribute at \nleast 50\\% of the flux in many of the strong lines, even when NGC 5548 is in its \nlowest state (Paper I). \n\nGiven the ionization parameters and densities of the two components from our \nstandard model, as well as the ionizing luminosity, the \ncharacteristic sizes (i.e., radii) for the two emitting regions are 1 pc for \nINNER, and 70 pc for OUTER. \n(Using the higher continuum luminosity given by the {\\it EUVE} point in Figure 2 \nwould increase these values by a factor of only $\\sqrt{1.30}$, or 1.14.)\nThus, the NLR of NGC 5548 is {\\it physically} \ncompact, since the size of OUTER is much smaller than typical values of 200 -- \n1000 pc determined for Seyfert 2 galaxies, using the same methods that we have \ndescribed in this paper (Kraemer \\& Harrington 1986; Kraemer et al. 1994).\nAlthough there have been reports of extended emission from NGC 5548 (Wilson et \nal. 1989), the contribution to the integrated emission line spectrum from this \nregion is small (Peterson et al. 1995); we estimated the contribution to the \nnarrow UV lines outside of a 1$''$ aperture (330 pc for H$_{0}$ $=$ 75 km \ns$^{-1}$ Mpc$^{-1}$) to be only $\\sim$20\\%. This is consistent with our finding \nthat the majority of the narrow emission must arise in a region with a \n``diameter'' of 140 pc.\n\nPogge (1989) and Schmitt \\& Kinney (1996) claim that the\n{\\it apparent} size of the NLR in a Seyfert 1 galaxy is typically much \nsmaller than that of a Seyfert 2 galaxy, although this may be due to a selection \neffect, since most of the Seyfert 2s in the {it HST} archive were selected on \nthe basis of their extended emission (Wilson 1997).\nNevertheless, it is clear that the majority of the Seyfert 1 galaxies in these \nstudies are apparently compact. This cannot be explained \nby viewing angle alone, since the opening angles \nof the presumed ionization cones are large, and in the simplest version of the \nunified model, the apparent extent of most Seyfert galaxies should be much \nlarger, even if viewed ``pole-on'' (Schmitt \\& Kinney 1996).\nOur results show that a possible explanation for the small apparent size of the \nNLRs in many Seyfert 1 galaxies is that they are truly (i.e., physically) \ncompact. Schmitt \\& Kinney explain this phenomenon as a result of\nthe orientation of the obscuring torus with respect\nto the plane of the galaxy. However, their model does not explain the dominance\nof the high ionization lines in NGC 5548 and many other Seyfert 1 galaxies, a \nfeature not generally seen in Seyfert 2 galaxies \n(Koski 1978; Shuder \\& Osterbrock 1981). Thus, if the unified model applies we \nmight expect that the narrow emission line spectrum of this Seyfert 1 galaxy \nwould resemble that of Seyfert 2 galaxies, and, further, we would expect \nto see a noticeable contribution to the spectrum from gas hundreds of parsecs \nfrom the central source. It is possible that the high ionization region in \nSeyfert 2 galaxies is obscured by a torus, but this does not explain the absence \nof a low ionization component in NGC 5548. Not only do our models show that the \nnarrow-line spectrum can be well fitted without such a component, but it is \nclear that the NLR of NGC 5548 is dominated by high ionization gas that must \nbe located close to the central source.\n\nA few Seyfert 1 galaxies do indeed show significant NLR emission at\nlarge distances from the central source. For example, {\\it HST} observations of \nNGC 4151 (Evans et al. 1993; Hutchings et al. 1997) reveal an array of knots and \nfilaments out to a kiloparsec that are almost certainly ionized by the central \ncontinuum. It may be that some Seyfert 1 galaxies, such as NGC 4151, have \nextended emission line regions and narrow lines in their spectra that closely\nresemble Seyfert 2 galaxies. Cohen (1983) studied the optical narrow\nline spectra of a group of Seyfert 1 galaxies and found some resemblance\nalthough, as a group, they appeared to be of somewhat higher ionization.\nCertainly, some of the galaxies in his study have spectra that are\nindistinguishable from those of type 2 Seyferts, but others, like NGC 5548, \nappear to be dominated by high ionization lines, a condition that appears to be \nrare among Seyfert 2 galaxies. It would be extremely interesting if one could \ndetermine if these high ionization objects are as compact as NGC 5548 appears to \nbe. Unfortunately, optical spectra alone are insufficient for detailed \nmodeling of the NLR, and thus far, accurate measurements of the narrow-line \nstrengths in the UV have only been obtained for NGC 5548\nand NGC 4151 (Ferland \\& Mushotzky 1982).\n\nAs we stated above, our finding that the inner component of gas is\noptically thin at the Lyman limit is based both on the absence of strong\nMg II emission and the relative strength of the high ionization lines.\nIf this is indeed true, not only for NGC 5548 but for other Seyfert 1\ngalaxies, it may be a clue to the origin of the inner narrow-line\ngas. Thin filaments or knots of reasonably high density ($\\sim$ 10$^{7}$ \ncm$^{-3}$) could be the result of outflow from the BLR, either as condensations\nin an expanding intercloud medium or as ``tails'' of BLR clouds, driven\nout by radiation pressure. If so, the small physical depths of the clouds\ninferred from our modeling ($\\sim$ 10$^{14}$ cm) may constrain the sizes of \ntheir BLR progenitors. Note that there has been some recent success in \nmodels of the BLR using a component of\noptically thin clouds (Shields, Ferland, \\& Peterson 1995).\nThere have been other studies suggesting that the \nclouds in the NLR are matter bounded or that some mix of radiation bounded and \nmatter bounded clouds exist (Wilson et al. 1997; Viegas-Aldrovandi 1988). Such a \nmix might give rise to the filamentary structure seen in some [O III] images of \nSeyfert galaxies rather than the typical molecular clouds found in spiral \ngalaxies, and may give a clue about the origin of the NLR gas.\n\nOne other point regarding the NLR gas is that it appears that there\nmay be some dust present within the outer components of the emitting\nregion. Dust is certainly able to exist in clouds at this proximity\nto the central source, with dust temperatures reaching a few hundred\nK (Kraemer \\& Harrington 1986). Nevertheless, the history of the dust\nin this gas is unknown. Our finding that the outer clouds\nmay not be purely radiation bounded would indicate that they are not\nsimply interstellar molecular clouds that have an outer shell ionized\nby the central source. Our finding that there is probably dust mixed in with\nthis gas will be important in determining the origin of this component.\n\nFinally, Moore et al. (1996) find a correlation of ionization potential of the \nnarrow emission lines with velocity width in NGC 5548. Combined with our finding \nthat, to first order, the ionization level decreases with distance, this \nindicates that the radial velocities decrease with increasing distance. This \ntrend is also seen in a more direct fashion in spatially resolved spectra of the \ninner NLR in NGC 4151 (Hutchings et al. 1997).\n\n\\section{Conclusions}\n\nWe have analyzed UV and optical spectra of the Seyfert 1 galaxy NGC 5548 that \nwere obtained when it was at an historical minimum. We were able to \nisolate the narrow emission line components (which do not vary in flux on short \ntime scales), due to the relative weakness of the more rapidly variable broad \nemission lines that are usually blended with the narrow components.\nWe have constructed photoionization models of the narrow-line region of this \ngalaxy, and are able to successfully match the observed dereddened ratios of a \nlarge number of emission lines to within the errors, with the \nexceptions noted in Section 3.3. Since we used the best direct observational \nevidence of the shape of the ionizing continuum, rather than making adjustments \nbased on fitting emission line ratios, the quality of this fit is particularly \nsatisfying. The fact that a good fit was obtained for the permitted emission \nlines, such as C~IV~$\\lambda$1549, and the forbidden lines, such as [Ne III] \n$\\lambda$3869, [O III] $\\lambda$5007, [O II] $\\lambda$3727, [N II] \n$\\lambda$6584, etc., indicates that the range of physical conditions assumed \nin these models is approximately correct.\n\nFrom our analysis and modeling of these spectra, we can make several statements \nregarding the physical conditions in the NLR of NGC 5548. First, it is clear \nthat the principal source of ionization in the NLR of NGC 5548 is the central \ncontinuum source. This conclusion is borne out by the quality of the fit to the \nemission line spectrum. The NLR covering factor is reasonably small (C $=$ \n0.11), so the NLR gas does not need to intercept much of the ionizing continuum \nto produce the observed emission lines fluxes. A second conclusion is that the \nhighly ionized gas in the inner part of the NLR appears to be optically thin at \nthe Lyman limit ($\\tau_{0}$ $\\approx$ 2.5), which yields constraints on the \nphysical depths of these clouds and may provide a clue to their origin. We have \nalso presented evidence for supersolar abundances in the inner portion of the \nNLR, and dust in the outer portion. \n\nThe most important conclusion that we have reached in this study is\nthat the NLR of NGC 5548 is physically compact, with the majority of emission \ncoming from a distance $\\leq$~70 pc from the nucleus. By contrast, the unified \nmodel of Seyfert galaxies suggests that the {\\it physical} dimensions of the NLR \nin Seyfert 1 and 2 galaxies should be similar. Additional studies of the type \nthat we have presented in this paper, particularly of other Seyfert 1 galaxies, \nare important for testing this aspect of the unified model.\n\n\\acknowledgments\n\nWe thank the referee, Andrew Wilson, for helpful comments and suggestions.\nWe are grateful to Fred Bruhweiler and Pat Harrington for helpful discussions on \nthe physical properties of iron and the availability of atomic data.\nS.B.K. and D.M.C. acknowledge support from NASA grant NAG~5-4103.\nA.V.F acknowledges support from NASA grant NAG~5-3556, and B.M.P. \nacknowledges support from NSF grant AST-9420080.\n\n\\clearpage\t\t\t \n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\\label{SecIntro}\nA crucial step in the history of General Relativity (GR) was Einstein's adoption of the principle of general covariance which states that the form of our physical laws should be independent of any choice of coordinate systems. The conceptual benefits of writing a theory in a coordinate-free way are immense. A generally covariant formulation of a theory has at least two major benefits: 1) it more clearly exposes the theory's geometric background structure, and 2) it thereby helps clarify our understanding of the theory's symmetries (i.e., its structure\/solution preserving transformations). It does both of these by disentangling the theory's substantive content from representational artifacts which arise in particular coordinate representations~\\cite{Pooley2015,Norton1993,EarmanJohn1989Weas}. Thus, general covariance is an indispensable tool for a modern understanding of spacetime theories. \n\nMotivated by quantum gravity, one may wish to extend these notions to quantum spacetime theories (whatever those are). Relatedly, one might want to extend these notions to discrete spacetime theories (i.e., lattice theories\\footnote{Given the results of this paper and of \\cite{DiscreteGenCovPart1}, calling these ``lattice theories'' can be misleading. This would be analogous to referring to continuum spacetime theories as ``coordinate theories''. As I will discuss, in both cases the coordinate systems\/lattice structure are merely representational artifacts and so do not deserve ``first billing'' so to speak. All lattice theories are best thought of as lattice-representable theories. Similarly, the term ``discrete spacetime theories'' ought to be here read as ``discretely-representable spacetime theories''. As discussed here (and in \\cite{DiscreteGenCovPart1}), the defining feature of such theories is that they have a finite density of degrees of freedom, see the work of Achim Kempf~\\cite{UnsharpKempf,Kempf2003,Kempf2004a,Kempf2004b,Kempf2006}}.). In \\cite{DiscreteGenCovPart1} I developed two analogs of general covariance for such discrete spacetime theories in a non-Lorentzian setting. The aim of this paper is to extend these results to a Lorentzian setting. Indeed, the analysis provided here is nearly identical to the one carried out in \\cite{DiscreteGenCovPart1}, although each paper is self-contained.\n\nIn either setting these discrete analogs of general covariance reveal that lattice structure is rather less like a fixed background structure or a fundamental part of some underlying manifold and rather more like a coordinate system, i.e., merely a representational artifact. Indeed, these discrete analogs are built upon a rich analogy between the lattice structures appearing in our discrete spacetime theories and the coordinate systems appearing in our continuum spacetime theories. \n\nThis paper is largely inspired by the brilliant work of mathematical physicist Achim Kempf~\\cite{Kempf_1997,UnsharpKempf,Kempf2000b,Kempf2003,Kempf2004a,Kempf2004b,Kempf2006,Martin2008,Kempf_2010,Kempf2013,Pye2015,Kempf2018} among others~\\cite{PyeThesis,Pye2022,BEH_2020}. A key feature present both here and in Kempf's work is the sampling property of bandlimited function revealed by the Nyquist-Shannon sampling theory~\\cite{GARCIA200263,SamplingTutorial,UnserM2000SyaS}. I review sampling theory in more detail in Sec.~\\ref{SecSamplingTheory}, but let me overview here. Bandlimited functions are those with have a limited extent in Fourier space (i.e., compact support). Bandlimited functions have the following sampling property: they can be exactly reconstructed knowing only the values that they take on any sufficiently dense sample lattice. What ``sufficiently dense'' means is fixed in terms of the size of the function's support in Fourier space.\n\nNyquist-Shannon sampling theory was first discovered in the context of information processing as a way of converting between analog and digital signals (i.e., between continuous and discrete information). Sampling theory found its first application in fundamental spacetime physics with Kempf's \\cite{Kempf_1997,UnsharpKempf}, ultimately leading to his thesis that ``Spacetime could be simultaneously continuous and discrete, in the same way that information can be'' \\cite{Kempf_2010}. Kempf's thoughts on these topics is the primary inspiration for this paper and deserves wider appreciation by the philosophy of physics community. For an overview of Kempf's works on this topic see~\\cite{Kempf2018}.\n\nMy thesis in \\cite{DiscreteGenCovPart1} is in broad agreement with Kempf's with one crucial alteration. I stress that the sampling property of bandlimited functions indicates that bandlimited physics can be simultaneously \\textit{represented as} continuous and discrete, (i.e., on a continuous or discrete spacetime). However, I further argue (both here and in \\cite{DiscreteGenCovPart1}) that when one investigates these two representations one finds substantial issue with taking the discrete representation as fundamental. These issues stem from the rich analogy between the lattice structures and coordinate systems mentioned above.\n\nThis analogy is supported here (and in \\cite{DiscreteGenCovPart1}) by the three lessons each of which tell against an intuitions one is likely to have regarding lattice structure. To motivate these (wrong) intuitions, consider the following situation.\n\nSuppose that after substantial empirical investigation of our micro-physical reality we find what appear to be ``lattice artifacts''. For instance, we may find ourselves restricted to only quarter rotation, or one-sixth rotation symmetries. Intuitively, this would suggest that the world is fundamentally set on a lattice of the kind shown in Fig.~\\ref{FigLat}, i.e., a square or hexagonal lattice.\n\n\\begin{figure}[t!]\n\\includegraphics[width=0.4\\textwidth]{PaperFigures\/FigLat.pdf}\n\\caption{Three of the lattice structures in space considered throughout this paper. The arrows indicate the indexing conventions for the lattice sites. Repeated in time, these give us something like a square 2D lattice, a cubic 3D lattice, and a hexagonal 3D lattice respectively. [\\textit{Reproduced with permission from} \\cite{DiscreteGenCovPart1}.]}\\label{FigLat}\n\\end{figure}\n\nSuppose that we have great predictive success when modeling the world as being set on (for instance) a square lattice with next-to-nearest neighbor interactions. Would this in any way prove that the world is fundamentally set on such a lattice? No, all this would prove is that the world can be \\textit{faithfully represented} on such a lattice with such interactions, at least empirically. Anything can be faithfully represented in any number of ways, this is just mathematics. Some extra-empirical work must be done to know if we should take the lattice structure appearing in this representation seriously. That is, we must ask which parts of the theory are substantive and which parts are merely representational? The discrete analogs of general covariance developed here (and in \\cite{DiscreteGenCovPart1}) answer this question: lattice structures are coordinate-like representational artifacts and so ultimately have no physical content.\n\nTo flesh out the contrary received position however, let us proceed without this analogy for the moment. We can ask: beyond merely appearing in our hypothetical empirically successful theory, what reason do we have to take the lattice structures which appear in this theory seriously? Well, intuitively the lattice structures appear to play a very substantial role in these theories, not merely a representational one. One likely has the following three interconnected first intuitions regarding the role that the lattice and lattice structure play in discrete spacetime theories:\n\\begin{itemize}\n\\item[1.] They restrict our possible symmetries. Taking the lattice structure to be a part of the theory's fixed background structure, our possible symmetries are limited to those which preserve this fixed structure. Intuitively a theory set on a square lattice can only have the symmetries of that lattice. Similarly for a hexagonal lattice, or even an unstructured lattice.\n\\item[2.] Differing lattice structures distinguishes our theories. Two theories with different lattice structures (e.g., square, hexagonal, irregular, etc.) cannot be identical. As suggested above they have different fixed background structures and as therefore have different symmetries.\n\\item[3.] The lattice is fundamentally ``baked-into'' the theory. Firstly, it is what the fundamental fields are defined over: they map lattice sites (and in \\cite{DiscreteGenCovPart1} times) into some value space. Secondly, the bare lattice is what the lattice structure structures. Thirdly, it is what limits us to discrete permutation symmetries in advance of further limitations from the lattice structure.\n\\end{itemize}\nThese intuitions will be fleshed out and made more concrete in Sec.~\\ref{SecKlein1}. However, as this paper demonstrates, each of the above intuitions are doubly wrong and overhasty.\n\nWhat goes wrong with the above intuitions is that we attempted to directly transplant our notions of background structure and symmetry from continuous to discrete spacetime theories. This is an incautious way to proceed and is apt to lead us astray. Recall that, as discussed above, our notions of background structure and symmetry are best understood in light of general covariance. It is only once we understand what is substantial and what is merely representational in our theories that we have any hope of properly understanding them. Therefore, we ought to instead first transplant a notion of general covariance into our discrete spacetime theories and then see what conclusions we are led to regarding the role that the lattice and lattice structure play in our discrete spacetime theories. This transplant has been done in a non-Lorentzian setting in \\cite{DiscreteGenCovPart1}. Here I extend these results to a Lorentzian setting.\n\nThis paper will teach us three lessons each of which negates one of the above intuitions about the role that lattice structure plays in discrete spacetime theories.\n\nFirstly, as I will show, taking any lattice structure seriously as a fixed background structure systematically under predicts the symmetries that discrete theories can and do have. Indeed, as I will show neither the bare lattice itself nor its lattice structure in any way restrict a theory's possible symmetries. In \\cite{DiscreteGenCovPart1}, for non-Lorentzian theories I have shown that there is no conceptual barrier to having a theory with continuous translation and rotation symmetries formulated on a discrete lattice. Indeed, in \\cite{DiscreteGenCovPart1} I presented a perfectly rotation invariant lattice theory. As I discuss in \\cite{DiscreteGenCovPart1}, this is analogous to the familiar fact that there is no conceptual barrier to having a continuum theory with rotational symmetry formulated on a Cartesian coordinate system. Here, I repeat this analysis in a Lorentzian context. In Sec.~\\ref{PerfectLorentz}, I present a perfectly Lorentzian lattice theory. \n\nSecondly, as I will show, discrete theories which are initially presented to us with very different lattice structures (i.e., square vs. hexagonal) may nonetheless turn out to be completely equivalent theories or to be overlapping parts of some larger theory. Moreover, given any discrete theory with some lattice structure we can always re-describe it using a different lattice structure. As I will discuss, this is analogous to the familiar fact that our continuum theories can be described in different coordinates, and moreover we can switch between these coordinate systems freely.\n\nThirdly, as I will show, in addition to being able to switch between lattice structures, we can also reformulate any discrete theory in such a way that it has no lattice structure whatsoever. Indeed, we can always do away with the lattice altogether. As I will discuss, this is analogous to the familiar fact that any continuum theory can be written in a generally covariant (i.e., coordinate-free) way.\n\nThese three lessons combine to give us a rich analogy between lattice structures and coordinate systems. It is from this rich analogy that the central claims of this paper follow. Namely, from this analogy it follows that the lattice structure supposedly underlying any discrete ``lattice'' theory has the same level of physical import as coordinates do, i.e., none at all. Thus, as I argued in \\cite{DiscreteGenCovPart1}, the world cannot be ``fundamentally set on a square lattice'' (or any other lattice) any more than it could be ``fundamentally set in a certain coordinate system''. Like coordinate systems, lattice structures are just not the sort of thing that can be fundamental; they are both thoroughly merely representational. Spacetime cannot be a lattice (even when it might be representable as such). Specifically, I claimed that properly understood, there are no such things as lattice-fundamental theories, rather there are only lattice-representable theories. This paper extends these conclusions to a Lorentzian context.\n\nOnce one begins thinking of lattices as a merely representational structure, a path opens for perfectly Lorentzian lattice theories. As proponents of causal set theory correctly point out, no single fixed spacetime lattice is Poincar\\'e invariant. This (apparently) spells big trouble for any lattice-based Lorentzian theories. They, however, avoid this issue by considering instead a random Poisson sprinkling of lattice points which does not pick out any preferred direction and hence does not explicitly break Poincar\\'e symmetry, at least on average. However, given the deflationary position this paper takes towards lattices, I claim there is no issue to be avoided. Like coordinate systems, lattice structures are just a representational tool for helping us express our theory. There is no need for the symmetries of our representational tools to latch onto the symmetries of the thing being represented. Cartesian coordinates are distorted under Lorentz boosts, but we can still use them to describe our Lorentzian theories without issue. The same is true of lattices. Indeed, in Sec.~\\ref{PerfectLorentz} I will present a perfectly Lorentzian lattice theory.\n\n\\subsection{Outline of the Paper}\nIn Sec.~\\ref{SecSevenKG}, I will introduce seven discrete Klein Gordon equations in an interpretation-neutral way and solve their dynamics. Then, in Sec.~\\ref{SecKlein1}, I will make a first attempt at interpreting these theories. I will (ultimately wrongly) identify their underlying manifold, locality properties, and symmetries. Among other issues, a central problem with this first attempt is that it takes the lattice itself to be the underlying spacetime manifold and thereby unequivocally cannot support continuous translation and rotation symmetries. This systematically under predicts the symmetries that these theories can and do have.\n\nIn Sec.~\\ref{SecKlein2}, I will provide a second attempt at interpreting these theories which fixes this issue (albeit in a slightly unsatisfying way). In particular, in this second attempt I deny that the lattice is the underlying spacetime manifold. Instead, I ``internalize'' it into the theory's value space. Fruitfully, this second interpretation does allow for continuous translation and rotation symmetries and even a (limited) Lorentz boost symmetry. However, the key move here of ``internalization'' has several unsatisfying consequences. For instance, the continuous symmetries we find here are all classified as internal (i.e, associated with the value space) whereas intuitively they ought to be external (i.e, associated with the manifold).\n\nWe thus will need a third attempt at interpreting these theories which externalizes these symmetries. Sec.~\\ref{SecExtPart1} - Sec.~\\ref{SecExtPart2} lay the groundwork for this third interpretation. In particular, they describe a principled way of 1) inventing a continuous spacetime manifold for our formerly discrete theories to live on and 2) embedding our theory's states\/dynamics onto this manifold as a new dynamical field. In the middle of this, in Sec.~\\ref{SecSamplingTheory}, I will provide an informal overview of the primary mathematical tools used in the latter half of this paper. Namely, I will review the basics of Nyquist-Shannon sampling theory and bandlimited functions. \n\nWith this groundwork complete, in Sec.~\\ref{SecKlein3} and Sec.~\\ref{SecKlein3Extra} I will provide a third attempt at interpreting these seven theories which fixes all issues arising in the previous two interpretations. For instance, like in my second attempt, this third interpretation can support continuous translation and rotation symmetries as well as a (limited) Lorentz boost symmetry. However, unlike the second attempt it realizes them as external symmetries (i.e., associated with the underlying manifold, not the theory's value space).\n\nIn Sec.~\\ref{SecDisGenCov}, I will review the lessons learned in these three attempts at interpretation. As I will discuss, the lessons learned combine to give us a rich analogy between lattice structures and coordinate systems. As I will discuss, there are actually two ways of fleshing out this analogy: one internal and one external. This section spells out these analogies in detail, each of which gives us a discrete analog of general covariance. I find reason to prefer the external notion, but either is likely to be fruitful for further investigation\/use. Sec.~\\ref{PerfectLorentz} provides us with a perfectly Lorentzian lattice theory as promised.\n\nFinally, in Sec. \\ref{SecConclusion} I will summarize the results of this paper, discuss its implications, and provide an outlook of future work.\n\nFor comments on what this means for the the dynamical vs geometrical spacetime debate~\\cite{EarmanJohn1989Weas,TwiceOver,BelotGordon2000GaM,Menon2019,BrownPooley1999,Nonentity,HuggettNick2006TRAo,StevensSyman2014Tdat,DoratoMauro2007RTbS,Norton2008,Pooley2013} see \\cite{DiscreteGenCovPart1}. Here I will focus on the implications this work has for quantum gravity especially causal set theory~\\cite{Surya2019}.\n\n\\section{Seven Discrete Klein Gordon Equations}\\label{SecSevenKG}\nIn this section I will introduce seven discrete Klein Gordon equations (KG1-KG7) in an interpretation-neutral way and solve their dynamics. These theories are all describable as being set on a lattice in both space and time. In each of these theories the lattice in space will simply repeat itself in time. I consider the following three cases for the lattice in space: a uniform 1D lattice, a square 2D lattice, and a hexagonal 2D lattice, see Fig.~\\ref{FigLat}. Repeated in time, these give us something like a square 2D lattice, a cubic 3D lattice, and a hexagonal 3D lattice respectively.\n\nAs harmful as these choices seem to be to Lorentz invariance (as well as continuous translation and rotation invariance) as I will show they ultimately pose no barrier to our theories having these symmetries. As I will argue, these choices of lattice are ultimately merely choices of representation which have absolutely nothing to do with the thing being represented. In particular, there is no need for our representational structure to have the same symmetries as the thing being represented. There is no issue with using Cartesian coordinates to describe a rotationally invariant state\/dynamics. I claim that analogously there is no issue with using a lattice to describe a state\/dynamics with continuous translation and rotation invariance and even Lorentz invariance. This claim has already been demonstrated in \\cite{DiscreteGenCovPart1} for states\/dynamics with continuous translation and rotation invariance. This paper extends this claim about lattice structures to Lorentz invariance as well.\n\n\\subsection{Introducing KG1-KG7}\nTo begin, let us consider the continuum Klein Gordon equations in $1+1$ and $2+1$ dimensions:\n\\begin{align}\\label{KG00}\n&\\text{\\bf Continuum Klein Gordon Eq. (KG00):}\\\\\n\\nonumber\n&\\partial_{t}^2\\varphi(t,x) = (\\partial_x^2-M^2) \\, \\varphi(t,x)\\\\\n\\label{KG0}\n&\\text{\\bf Continuum Klein Gordon Eq. (KG0):}\\\\\n\\nonumber\n&\\partial_{t}^2\\varphi(t,x,y) = (\\partial_x^2+\\partial_y^2-M^2) \\, \\varphi(t,x,y)\n\\end{align}\nwith some mass $M\\geq0$. For a generally covariant (i.e., coordinate-free) view of these theories, see Sec.~\\ref{SecFullGenCov}. \n\nFor our first three discrete Klein Gordon theories, let us consider the theories with only nearest-neighbor (N.N.) interactions on the above discussed lattices which best approximate KG0 and KG00. Namely:\n\\begin{align}\\label{KG1Long}\n&\\text{\\bf 1D N.N. Klein Gordon Eq. (KG1):}\\\\\n\\nonumber\n&[\\phi_{j-1,n}-2\\phi_{j,n}+\\phi_{j+1,n}]\\\\\n\\nonumber\n&\\quad=[\\phi_{j,n-1}-2\\phi_{j,n}+\\phi_{j,n+1}]\n-\\mu^2 \\phi_{j,n}\\\\\n&\\text{\\bf Square N.N. Klein Gordon Eq. (KG4):}\\label{KG4Long}\\\\\n\\nonumber\n&[\\phi_{j-1,n,m}-2\\phi_{j,n,m}+\\phi_{j+1,n,m}]\\\\\n\\nonumber\n&\\quad = [ \\ \\phi_{j,n-1,m}-2\\phi_{j,n,m}+\\phi_{j,n+1,m}\\\\\n\\nonumber\n&\\quad \\ \\ +\\phi_{j,n,m-1}-2\\phi_{j,n,m}+\\phi_{j,n,m+1}]\n-\\mu^2 \\phi_{j,n,m}\\\\\n&\\text{\\bf Hexagonal N.N. Klein Gordon Eq. (KG5):}\\label{KG5Long}\\\\\n\\nonumber\n&[\\phi_{j-1,n,m}-2\\phi_{j,n,m}+\\phi_{j+1,n,m}]\\\\\n\\nonumber\n&\\quad = \\frac{2}{3}[ \\ \\phi_{j,n-1,m}-2\\phi_{j,n,m}+\\phi_{j,n+1,m}\\\\\n\\nonumber\n&\\qquad \\quad +\\phi_{j,n,m-1}\\!-\\!2\\phi_{j,n,m}\\!+\\!\\phi_{j,n,m+1}\\\\\n\\nonumber\n&\\qquad \\quad +\\phi_{j,n+1,m-1}-2\\phi_{j,n,m}+\\phi_{j,n-1,m+1}]\n-\\mu^2 \\phi_{j,n,m}\n\\end{align}\nwith $j\\in\\mathbb{Z}$ indexing time and $n\\in\\mathbb{Z}$ and $m\\in\\mathbb{Z}$ indexing space. See Fig.~\\ref{FigLat} for the indexing convention. Here $\\mu\\in\\mathbb{R}$ is a dimensionless number playing the role of the field's mass. The terms in square brackets in the above expressions are the best possible approximations of the second derivative on each lattice which make use of only nearest neighbor interactions. These theories are named KG1, KG4, and KG5 in correspondence with the discrete heat equations considered in \\cite{DiscreteGenCovPart1} and in anticipation of their further treatment later in this section. \n\nThis section has promised to introduce these theories in an interpretation neutral way. As such, some of the above discussion needs to be hedged. In particular, in introducing these theories I have made casual comparison between parts of these theories' dynamical equations and various approximations of the second derivative. While, as I will discuss, such comparisons can be made, to do so immediately is unearned. It comes dangerously close to imagining the spacetime lattices discussed above as being embedded in a continuous manifold. This may be something we want to do later (see Sec.~\\ref{SecExtPart1}), but it is a non-trivial interpretational move which ought not be done so casually. \n\nCrucially, in this paper I will begin by analyzing these theories as discrete-native theories. As such, it's important to think of these discrete spacetime theories as self-sufficient theories in their own right. We must not begin by thinking of them as various discretizations or bandlimitations of the continuum theories. While, as I will discuss, these discrete theories have some notable relationships to various continuum theories it is important to resist any temptation to see these continuum theories as ``where they came from''. Rather, let us pretend these theories ``came from nowhere'' and let us see what sense we can make of them.\n\nAnother bit of hedging: in introducing the above three theories I casually associated them with the lattice structures shown in Fig.~\\ref{FigLat} (each repeated in time). Making such associations ab initio is unwarranted. While we may eventually associate these theories with those lattice structures we cannot do so immediately. Such an association would need to be made following careful consideration of the dynamics. (Such an exercise is carried out in Sec.~\\ref{SecKlein1}.) Beginning here in an interpretation-neutral way these theories ought to be seen as being defined over a completely unstructured lattice. \n\nI will reflect this concern in my notation as follows. The labels for the lattice sites are presently too structured (e.g., $(j,n)\\in\\mathbb{Z}^2$ and $(j,n,m)\\in\\mathbb{Z}^3$). Instead we ought to think of the lattice sites as having labels $\\ell\\in L$ for some set $L$. Crucially, at this point the set of labels for the lattice sites, $L$, is just that, an unstructured set.\n\nUp to isomorphism (here, generic bijections, i.e. generic relabelings), sets are uniquely specified by their cardinality. The set of labels for the lattice sites is here countable, $\\ell\\in L\\cong\\mathbb{Z}\\cong\\mathbb{Z}^2\\cong\\mathbb{Z}^3$. Reframed this way the above discussed theories each consider the same discrete variables $\\phi_\\ell\\in\\mathbb{R}$. In particular, KG1 considers variables $\\phi_\\ell$ which under some convenient relabeling of the lattice sites, $\\ell\\in L\\mapsto (j,n)\\in\\mathbb{Z}^2$, satisfies Eq.~\\eqref{KG1Long}. Similarly, KG4 and KG5 consider variables $\\phi_\\ell$ which under some convenient relabeling of the lattice sites, $\\ell\\in L\\mapsto (j,n,m)\\in\\mathbb{Z}^3$, satisfy Eq.~\\eqref{KG4Long} and Eq.~\\eqref{KG5Long} respectively.\n\nIt's important to stress that the mere existence of these convenient relabelings by itself has no interpretative force. The fact that our labels $(j,n)\\in\\mathbb{Z}^2$ and $(j,n,m)\\in\\mathbb{Z}^3$ in some sense form a square 2D lattice and cubic 3D lattice in no way forces us to think of $L$ as being structured in this way (indeed, we might later like to think of $L$ as a hexagonal 3D lattice). In particular, the fact that these labels are in a sense equidistant from each other does not force us to think of the lattice sites as being equidistant from each other. Nor are we forced to think that ``the distance between lattice sites'' to be meaningful at all. Dynamical considerations may later push us in this direction, but the mere convenience of this labeling should not.\n\nI have above introduced three out of seven discrete Klein Gordon theories. In order to introduce the other four theories, it is convenient (but not necessary) to first reformulate things. In particular, let us reorganize the $\\phi_\\ell$ variables into a vector, namely,\n\\begin{align}\\label{PhiDef}\n\\bm{\\Phi}= \\sum_{\\ell\\in L} \\phi_\\ell \\, \\bm{b}_\\ell. \n\\end{align}\nwhere $\\bm{b}_\\ell$ is a linearly-independent basis vector for each $\\ell\\in L$ and \\mbox{$\\bm{\\Phi}$} is a vector in the vector space \\mbox{$\\mathbb{R}^L\\coloneqq\\text{span}(\\{\\bm{b}_\\ell\\}_{\\ell\\in L})$}. For later reference, it should be noted that $\\phi_\\ell$ is also a vector in a vector space: namely, $F_L$ the space of functions $f:L\\to\\mathbb{R}$. Note that Eq.~\\eqref{PhiDef} is an vector space isomorphism between these vector spaces, $\\mathbb{R}^L\\cong F_L$. Everything which follows concerning $\\bm{\\Phi}\\in\\mathbb{R}^L$ has an isomorphic description in terms of $\\phi_\\ell\\in F_L$.\n\nRecall that for KG1 the lattice sites $\\ell\\in L$ have a convenient relabeling in terms of two integer indices, \\mbox{$\\ell\\in L\\mapsto (j,n)\\in\\mathbb{Z}^2$}. We can use this relabeling to grant the vector space a tensor product structure as \\mbox{$\\mathbb{R}^L\\mapsto\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$} by taking \\mbox{$\\bm{b}_\\ell\\mapsto \\bm{e}_j\\otimes\\bm{e}_n$} where \n\\begin{align}\n\\bm{e}_m = (\\dots,0,0,1,0,0,\\dots)^\\intercal\\in\\mathbb{R}^\\mathbb{Z}\n\\end{align}\nwith the 1 in the $m^\\text{th}$ position. Under this restructuring of KG1 we have, \n\\begin{align}\\label{PhiVec1}\n\\bm{\\Phi}= \\sum_{j,n\\in\\mathbb{Z}} \\phi_{j,n} \\, \\bm{e}_j\\otimes\\bm{e}_n. \n\\end{align}\nIn these terms the dynamics of KG1 is given by,\n\\begin{align}\n\\label{DKG1}\n&\\text{\\bf Klein Gordon Equation 1 (KG1):}\\\\\n\\nonumber\n&\\Delta_{(1),\\text{j}}^2\\,\\bm{\\Phi}=\\Delta_{(1),\\text{n}}^2 \\, \\bm{\\Phi}\n-\\mu^2\\bm{\\Phi}\n\\end{align}\nwhere the notation $A_\\text{j}\\coloneqq A\\otimes\\openone$ and $A_\\text{n}\\coloneqq\\openone\\otimes A$ mean $A$ acts only on the first or second tensor space respectively. The linear operator $\\Delta_{(1)}^2$ appearing twice in the above expression is the following bi-infinite Toeplitz matrix:\n\\begin{align}\\label{Delta12}\n\\Delta_{(1)}^2=\\{\\Delta^+,\\Delta^-\\}\n&=\\text{Toeplitz}(1,\\,-2,\\,1)\\\\\n\\nonumber\n\\Delta^+&=\\text{Toeplitz}(0,-1,\\,1)\\\\\n\\nonumber\n\\Delta^-&=\\text{Toeplitz}(-1,\\,1,\\,0)\n\\end{align}\nwhere the curly brackets indicate the anticommutator, $\\{A,B\\}= \\frac{1}{2}(A\\,B + B\\,A)$. Recall that Toeplitz matrices are so called diagonal-constant matrices with \\mbox{$[A]_{i,j}=[A]_{i+1,j+1}$}. Thus, the values in the above expression give the matrix's values on either side of the main diagonal.\n\nAlthough above I warned about thinking in terms of derivative approximations prematurely, a few comments are here warranted. Note that $\\Delta^+$ is associated with the forward derivative approximation, $\\Delta^-$ is be associated with the backwards derivative approximation, and $\\Delta_{(1)}^2$ is associated with the nearest neighbor second derivative approximation,\n\\begin{align}\n\\nonumber\n\\Delta_{(1)}^2\/\\epsilon^2:\\ \\partial_x^2 f(x)\n&\\approx\\frac{f(x+\\epsilon)-2 f(x)+f(x-\\epsilon)}{\\epsilon^2}.\n\\end{align}\nAs stressed above, we ought to be cautious not to lean too heavily on these relationships when interpreting these discrete theories. \n\nIn addition to KG1, I will also consider two more theories with ``improved derivative approximations''. Namely,\n\\begin{align}\n\\label{DKG2}\n&\\text{\\bf Klein Gordon Equation 2 (KG2):}\\\\\n\\nonumber\n&\\Delta_{(2),\\text{j}}^2\\,\\bm{\\Phi}=\\Delta_{(2),\\text{n}}^2 \\ \\bm{\\Phi}-\\mu^2 \\, \\bm{\\Phi}\\\\\n\\label{DKG3}\n&\\text{\\bf Klein Gordon Equation 3 (KG3):}\\\\\n\\nonumber\n&D_{\\text{j}}^2\\,\\bm{\\Phi}=D_{\\text{n}}^2 \\ \\bm{\\Phi}-\\mu^2 \\, \\bm{\\Phi}\n\\end{align}\nwhere\n\\begin{align}\\label{BigToeplitz}\n\\Delta_{(2)}^2&=\\text{Toeplitz}(\\frac{-1}{12},\\,\\frac{4}{3},\\,\\frac{-5}{2},\\,\\frac{4}{3},\\,\\frac{-1}{12})\\\\\n\\nonumber\nD&=\\text{Toeplitz}(\\dots,\\!\\frac{-1}{5},\\!\\frac{1}{4},\\!\\frac{-1}{3},\\!\\frac{1}{2},\\!-1,\\!0,\\!1,\\!\\frac{-1}{2},\\!\\frac{1}{3},\\!\\frac{-1}{4},\\!\\frac{1}{5},\\!\\dots)\\\\\n\\nonumber\nD^2&=\\text{Toeplitz}(\\dots,\\!\\frac{-2}{16},\\!\\frac{2}{9},\\!\\frac{-2}{4},\\!\\frac{2}{1},\\!\\frac{-2\\pi^2}{6},\\!\\frac{2}{1},\\!\\frac{-2}{4},\\!\\frac{2}{9},\\!\\frac{-2}{16},\\!\\dots).\n\\end{align}\nNote that $\\Delta_{(2)}^2$ is related to the next-to-nearest-neighbor approximation to the second derivative. Obviously, the longer range we make our derivative approximations the more accurate they can be. The infinite-range operator $D$ (and its square $D^2$) in some sense are the best discrete approximations to the derivative (and second derivative) possible. The defining property of $D$ is that it is diagonal in the (discrete) Fourier basis with spectrum,\n\\begin{align}\\label{LambdaD}\n\\lambda_D(k)=-\\mathrm{i}\\,\\underline{k} \n\\end{align}\nwhere $\\underline{k}=k$ for $k\\in[-\\pi,\\pi]$ repeating itself cyclically with period $2\\pi$ outside of this region. This is in tight connection with the continuum derivative operator $\\partial_x$ which is diagonal in the (continuum) Fourier basis with spectrum \\mbox{$\\lambda_{\\partial_x}(k)=-\\mathrm{i} \\, k$} for $k\\in[-\\infty,\\infty]$. \n\nAlternatively, one can construct $D^2$ in the following way: generalize $\\Delta_{(1)}^2$ and $\\Delta_{(2)}^2$ to $\\Delta_{(n)}^2$ namely the best second derivative approximation which considers up to $n^\\text{th}$ neighbors to either side. Taking the limit $n\\to\\infty$ gives $D^2=\\lim_{n\\to\\infty}\\Delta_{(n)}^2$. Other aspects of $D$ will be discussed in Sec.~\\ref{SecSamplingTheory} (including its related derivative approximation Eq.~\\eqref{ExactDerivative}) but enough has been said for now.\n\nWhile these connections to derivative approximations allow us to export some intuitions from the continuum theories into these discrete theories, we must resist this (at least for now). In particular, I should stress again that we should not be thinking of any of KG1, KG2 and KG3 as coming from the continuum theory under some approximation of the derivative.\n\nLet's next reformulate KG4 and KG5 in terms of \\mbox{$\\bm{\\Phi}\\in\\mathbb{R}^L$}. In these cases we have a convenient relabeling of the lattice sites in terms of three integer indices, $\\ell\\mapsto (j,n,m)$. As before we can use this relabeling to grant the vector space a tensor product structure as $\\mathbb{R}^L\\mapsto\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$ by taking by taking \\mbox{$\\bm{b}_\\ell\\mapsto \\bm{e}_j\\otimes\\bm{e}_n\\otimes\\bm{e}_m$}. Under this restructuring we have,\n\\begin{align}\\label{PhiVec2}\n\\bm{\\Phi}= \\sum_{j,n,m\\in\\mathbb{Z}} \\phi_{j,n,m} \\, \\bm{e}_j\\otimes\\bm{e}_n\\otimes\\bm{e}_m. \n\\end{align}\nIn these terms the dynamics of KG4 given above (namely, Eq.~\\eqref{KG4Long}) is now given by,\n\\begin{align}\\label{DKG4}\n&\\text{\\bf Klein Gordon Equation 4 (KG4):}\\\\\n\\nonumber\n&\\Delta_{(1),\\text{j}}^2\\,\\bm{\\Phi}=(\\Delta_{(1),\\text{n}}^2+\\Delta_{(1),\\text{m}}^2) \\ \\bm{\\Phi}-\\mu^2 \\, \\bm{\\Phi}\n\\end{align}\nA similar treatment of the dynamics of KG5 (namely, Eq.~\\eqref{KG5Long}) gives us,\n\\begin{align}\\label{DKG5}\n&\\text{\\bf Klein Gordon Equation 5 (KG5):}\\\\\n\\nonumber\n&\\Delta_{(1),\\text{j}}^2\\,\\bm{\\Phi} =\\frac{2}{3} \\, \\Big[\\Delta^2_{(1),\\text{n}}+ \\Delta^2_{(1),\\text{m}}\\\\\n\\nonumber\n&\\qquad\\qquad\\quad \\ +\\big\\{\\Delta^+_\\text{m}-\\Delta^+_\\text{n},\\Delta^-_\\text{m}-\\Delta^-_\\text{n}\\big\\}\\Big] \\, \\bm{\\Phi}-\\mu^2 \\, \\bm{\\Phi}\n\\end{align}\nWhile the third term in the square brackets looks complicated, it is just the analog of $\\Delta^2_{(1),\\text{n}}$ and $\\Delta^2_{(1),\\text{m}}$ but in the $m-n$ direction. See Eq.~\\eqref{Delta12}.\n\nFinally, in addition to KG4 and KG5 I consider the following two theories:\n\\begin{align}\n\\label{DKG6}\n&\\text{\\bf Klein Gordon Equation 6 (KG6):}\\\\\n\\nonumber\n&D_{\\text{j}}^2\\,\\bm{\\Phi}=(D_{\\text{n}}^2+D_{\\text{m}}^2) \\ \\bm{\\Phi}-\\mu^2 \\, \\bm{\\Phi}\\\\\n\\label{DKG7}\n&\\text{\\bf Klein Gordon Equation 7 (KG7):}\\\\\n\\nonumber\n&D_{\\text{j}}^2\\,\\bm{\\Phi}= \\frac{2}{3}\\left(D^2_\\text{n}+D^2_\\text{m}+(D_\\text{m}-D_\\text{n})^2\\right) \\, \\bm{\\Phi}-\\mu^2 \\, \\bm{\\Phi}\n\\end{align}\nwhich resemble KG4 and KG5 but which make use of an infinite range coupling between lattice sites. Having introduced these seven theories, let us next solve their dynamics.\n\n\\subsection{Solving Their Dynamics}\nConveniently, each of KG1-KG7 admit planewave solutions. Moreover, in each case these planewave solutions form a complete basis of solutions.\n\nConsidering first KG1-KG3 we have solutions of the form,\n\\begin{align}\\label{PlaneWave123}\n\\phi_{j,n}(\\omega,k)=e^{-\\mathrm{i}\\, \\omega \\, j - \\mathrm{i}\\, k \\, n}.\n\\end{align}\nwith $\\omega,k\\in\\mathbb{R}^2$. It should be noted however, that outside of the range $\\omega,k\\in[-\\pi,\\pi]$ these planewaves repeat themselves with period $2\\pi$ due to Euler's identity, $\\exp(2\\pi\\mathrm{i})=1$. In terms of $\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$ these planewaves are:\n\\begin{align}\n\\bm{\\Phi}(\\omega,k)=\\sum_{j,n\\in\\mathbb{Z}} \\phi_{j,n}(\\omega,k) \\, \\bm{e}_j\\otimes\\bm{e}_n. \n\\end{align}\nFrom this planewave basis we can recover the $\\bm{e}_j\\otimes\\bm{e}_n$ basis as:\n\\begin{align}\n\\bm{e}_j\\otimes\\bm{e}_n\n&=\\frac{1}{(2\\pi)^2}\\int\\!\\!\\!\\!\\int_{-\\pi}^\\pi e^{\\mathrm{i}\\,\\omega \\,j+\\mathrm{i}\\, k\\, n}\\, \\bm{\\Phi}(\\omega,k)\\,\\d\\omega\\d k. \n\\end{align}\nThese planewaves are only a solution if $\\omega$ and $k$ satisfy the theory's dispersion relation which can be straight-forwardly calculated from the theory's dynamics:\n\\begin{align}\n\\text{KG1:}& \\quad \\!\\! 2-2\\,\\cos(\\omega)= \\mu^2+2-2\\,\\cos(k)\\\\\n\\text{KG2:}& \\quad \\!\\! \\frac{1}{6}\\,(\\cos(2\\,\\omega)-16\\,\\cos(\\omega)+15)\\\\\n\\nonumber\n&= \\mu^2+\\frac{1}{6}\\,(\\cos(2\\,k)-16\\,\\cos(k)+15)\\\\\n\\text{KG3:}& \\quad \\!\\! \\underline{\\omega}^2= \\mu^2+\\underline{k}^2\n\\end{align}\nwhere $\\underline{\\omega}=\\omega$ and $\\underline{k}=k$ for $\\omega,k\\in[-\\pi,\\pi]$ repeating themselves cyclically with period $2\\pi$ outside of this region. Note that the dispersion relation for KG3 follows from Eq.~\\eqref{LambdaD}, essentially from the definition of $D$.\n\n\\begin{figure}[t!]\n\\includegraphics[width=0.45\\textwidth]{PaperFigures\/FigKleinDisp.pdf}\n\\caption{The dispersion relations for the planewave solutions to the discrete Klein Gordon equations are plotted as a function of wavenumber for KG1, KG2 and KG3 with $\\mu=1.25$.}\\label{FigKleinDisp}\n\\end{figure}\n\nFig.~\\ref{FigKleinDisp} shows these dispersion relations restricted to the region $\\omega,k\\in[-\\pi,\\pi]$ with a field mass of $\\mu=1.25$. Qualitatively, KG1-KG3 all seem to agree with each other at low wavenumbers. They appear to mostly differ with respects to the rate at which high wavenumber planewaves oscillate. Let's investigate how these theories behave for planewaves with periods and wavelengths which span many lattice sites, that is with $\\vert\\omega\\vert,\\vert k\\vert\\ll\\pi$.\n\\begin{align}\n\\nonumber\n\\text{KG1:}& \\quad \\!\\! \\omega^2= \\mu^2+k^2+\\frac{\\omega^4-k^4}{12}+\\mathcal{O}(\\omega^6)+\\mathcal{O}(k^6)\\\\\n\\nonumber\n\\text{KG2:}& \\quad \\!\\! \\omega^2= \\mu^2+k^2-\\frac{\\omega^6-k^6}{90}+\\mathcal{O}(\\omega^8)+\\mathcal{O}(k^8)\\\\\n\\text{KG3:}& \\quad \\!\\! \\omega^2 = \\mu^2+k^2.\n\\end{align}\nNote that the dispersion relation for KG3 exactly matches that of the continuum theory, not only within this regime but for all $\\omega,k\\in[-\\pi,\\pi]$. In the $\\vert\\omega\\vert,\\vert k\\vert\\ll\\pi$ regime, KG2 gives a better approximation of the continuum theory than KG1 does. This is due to its longer range coupling giving a better approximation of the derivative. \n\nIf we consider only solutions with all or most of their planewave support with $\\vert\\omega\\vert,\\vert k\\vert\\ll\\pi$, we have an approximate one-to-one correspondence between the solutions to these theories. This is roughly why each of these theories have the same continuum limit, namely KG00 defined above. In terms of the rate at which these theories converge to the continuum theory in the continuum limit, one can expect KG3 to outpace KG2 which outpaces KG1. (As I discussed in \\cite{DiscreteGenCovPart1}, this is in a way counter-intuitive: why does the most non-local discrete theory give the best approximation of our perfectly local continuum theory?)\n\nHowever, while interesting in their own right, these relationships with the continuum theory are not directly helpful in helping us understand KG1-KG3 in their own terms as discrete-native theories.\n\nMoving on to KG4-KG7, their planewave solutions are of the form, \n\\begin{align}\n\\phi_{j,n,m}(\\omega,k_1,k_2)=e^{-\\mathrm{i}\\,\\omega \\,j-\\mathrm{i}\\, k_1\\, n-\\mathrm{i}\\, k_2\\, m}\n\\end{align}\nwith $\\omega,k_1,k_2\\in\\mathbb{R}^2$. Again, it should be noted however, that outside of the range \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$} these planewaves repeat themselves with period $2\\pi$ due to Euler's identity, $\\exp(2\\pi\\mathrm{i})=1$. In terms of \\mbox{$\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$} these planewaves are:\n\\begin{align}\n\\nonumber\n\\bm{\\Phi}(\\omega,k_1,k_2)=\\sum_{j,n,m\\in\\mathbb{Z}} \\phi_{j,n,m}(\\omega,k_1,k_2) \\, \\bm{e}_j\\otimes\\bm{e}_n\\otimes\\bm{e}_m. \n\\end{align}\nFrom this planewave basis we can recover the $\\bm{e}_j\\otimes\\bm{e}_n\\otimes\\bm{e}_m$ basis as:\n\\begin{align}\n&\\bm{e}_j\\otimes\\bm{e}_n\\otimes\\bm{e}_m\\\\\n\\nonumber\n&=\\frac{1}{(2\\pi)^3}\\int\\!\\!\\!\\!\\int\\!\\!\\!\\!\\int_{-\\pi}^\\pi e^{\\mathrm{i}\\,\\omega \\,j+\\mathrm{i}\\, k_1 n+\\mathrm{i}\\, k_2 m}\\,\\bm{\\Phi}(\\omega,k_1,k_2)\\,\\d\\omega\\d k_1 \\d k_2. \n\\end{align}\n\nThe dispersion relation for each of these theories is given by:\n\\begin{align}\n\\text{KG4:}& \\ 2-2\\cos(\\omega)= \\mu^2\\!+\\!4\\!-\\!2\\cos(k_1)\\!-\\!2\\cos(k_2)\\\\\n\\nonumber\n\\text{KG5:}& \\ 2-2\\,\\cos(\\omega)= \\mu^2 \\\\\n\\nonumber\n&+\\frac{4}{3}\\big[3-\\cos(k_1)-\\cos(k_1)-\\cos(k_2-k_1)\\big]\\\\\n\\nonumber\n\\text{KG6:}& \\ \\underline{\\omega}^2= \\mu^2 +\\underline{k_1}^2 +\\underline{k_2}^2\\\\\n\\nonumber\n\\text{KG7:}& \\ \\underline{\\omega}^2= \\mu^2 +\\frac{2}{3}\\big[\\underline{k_1}^2 +\\underline{k_2}^2 +(\\underline{k_2}-\\underline{k_1})^2\\big].\n\\end{align}\nNote that the dispersion relation for KG6 and KG7 follow from Eq.~\\eqref{LambdaD}, essentially from the definition of $D$.\n\nUnlike KG1-KG3, these theories do not all agree with each other in the small $\\omega,k_1,k_2$ regime. KG4 and KG6 agree that for $\\vert\\omega\\vert, \\vert k_1\\vert, \\vert k_2\\vert \\ll \\pi$ we have \\mbox{$\\omega^2= \\mu^2 +k_1^2 +k_2^2$}. Moreover, KG5 and KG7 agree with each other in this regime, but not with KG4 and KG6. Do we have two different results in the continuum limit here?\n\nCloser examination reveals that we do not. The key to realizing this is to note that under the transformation, \n\\begin{align}\\label{SkewKG7KG6}\n\\omega\\mapsto \\omega,\\quad\nk_1\\mapsto k_1,\\quad\nk_2\\mapsto \\frac{1}{2}\\, k_1+\\frac{\\sqrt{3}}{2}\\,k_2.\n\\end{align}\nwe have the dispersion relation for KG7 mapping exactly onto the one for KG6. The inverse of this map is\n\\begin{align}\\label{SkewKG6KG7}\n\\omega\\mapsto \\omega,\\quad\nk_1\\mapsto k_1,\\quad\nk_2\\mapsto \\frac{2k_2-k_1}{\\sqrt{3}}.\n\\end{align}\nTechnically, when acting on the planewaves $\\bm{\\Phi}(\\omega,k_1,k_2)$ these transformations are only each other's inverses when we have \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$} both before and after the transformation. This is due to the $2\\pi$ periodicity of these planewaves. Fortunately however, all of KG7's planewave solutions with \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$} remain in this region after applying Eq.~\\eqref{SkewKG7KG6}. The same is true of KG6: its planewave solutions with \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$} also remain in this region after applying Eq.~\\eqref{SkewKG6KG7}. As I will soon discuss, this means we have an exact one-to-one correspondence between KG6 and KG7's solutions (much ado will be made about this later.) Applying this transformation to KG5 does not map it onto KG4, but it does bring their $\\vert\\omega\\vert,\\vert k_1\\vert,\\vert k_2\\vert\\ll\\pi$ behavior into agreement. \n\nThus, if we consider only solutions with all or most of their planewave support with $\\vert\\omega\\vert,\\vert k_1\\vert,\\vert k_2\\vert\\ll\\pi$ (or the appropriately transformed regime for KG5 and KG7) then we have an approximate one-to-one correspondence between the solutions to these theories. Within this regime we can define their common continuum limit, KG0. Repeating our analysis of the convergence rates of KG1-KG3 here, we expect KG6 and KG7 to converge in the continuum limit faster than KG4 and KG5 do.\n\n\\vspace{0.25cm}\n\nThis paper will make three attempts at interpreting these seven discrete theories. Allow me to identify in advance three important points of comparison between these interpretations. \n\nThe first important point of comparison is what sense they make of these different convergence rates in the continuum limit. As discussed above, in terms of this convergence rate we expect \\mbox{$\\text{KG3}>\\text{KG2}>\\text{KG1}$} and similarly \\mbox{$\\text{KG6}, \\text{KG7} > \\text{KG4}, \\text{KG5}$} with higher rated theories converging more quickly. This is in tension with our intuitive sense of locality for these theories: judging locality by the number of lattice sites coupled together we have\n\\mbox{$\\text{KG1} > \\text{KG2} > \\text{KG3}$} with higher rated theories being more local and similarly\n\\mbox{$\\text{KG4},\\text{KG5} > \\text{KG6},\\text{KG7}$}. How is it that our most non-local theories are somehow the nearest to our perfectly local continuum theory?\n\nRegarding how these three interpretations deal with this tension, not much changes between the discrete heat equations considered in \\cite{DiscreteGenCovPart1} and the discrete Klein Gordon equations considered here. As such, I will leave any detailed discussion of this issue to \\cite{DiscreteGenCovPart1} and direct the interested reader there. Roughly, the second and third interpretations deal with this tension by negating or reversing all of the above intuitive locality judgements.\n\nA second important point of comparison between these three interpretations will be what sense they make of the above-noted exact one-to-one correspondence between KG6 and KG7's solutions. (More will be said about this in Sec.~\\ref{SecKlein2}.) It is important to note that the mere existence of such a one-to-one correspondence does not automatically mean that these theories are identical or even equivalent; All it means technically is that their solution spaces have the same cardinality. As I will discuss, some of the coming interpretations recognize KG6 and KG7 as being equivalent whereas others do not.\n\nA third important point of comparison between the coming interpretations will be what sense they make of these theories having continuous symmetries. For instance, the dispersion relation for KG6 appears to be in some sense rotation invariant and even Lorentz invariant (at least in Fourier space and staying inside of the region \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$}). In a sense, KG7 might have these symmetries too: given the above-noted one-to-one correspondence between the solutions of KG6 and KG7, there may be some (skewed) sense in which KG7 is rotation invariant and Lorentz invariant as well. All of this will be made precise later on. As I will discuss, some interpretations consider KG6 and KG7 to have a rotation symmetry and even (limited) Lorentz boost symmetries whereas others do not. As I will discuss in Sec.~\\ref{PerfectLorentz}, KG6 and KG7 can be seen as representationally-limited parts of a larger perfectly Lorentzian lattice theory.\n\nHaving introduced these theories and solved their dynamics in an interpretation-neutral way. We can now make a first (ultimately misled) attempt at interpreting them.\n\n\\section{A First Attempt at Interpreting KG1-KG7}\\label{SecKlein1}\nNow that we have introduced these seven discrete theories and solved their dynamics, let's get on to interpreting them. Let us begin by following our first intuitions and analyze these seven discrete theories concerning their underlying manifold, locality properties and symmetries. Ultimately however, as I will discuss later, much of the following is misled and will need to be revisited and revised later. Luckily, retracing where we went wrong here will be instructive later.\n\nLet's start by taking the initial formulation of the above theories in terms of $\\phi_\\ell$ seriously, i.e. Eq.~\\eqref{KG1Long}, Eq.~\\eqref{KG4Long} and Eq.~\\eqref{KG5Long}. Taken literally as written, what are these theories about? Intuitively these theories are about a field $\\phi_\\ell$ which maps lattice sites ($\\ell\\in L\\cong\\mathbb{Z}\\cong\\mathbb{Z}^2\\cong\\mathbb{Z}^3$) into field amplitudes ($\\phi_\\ell\\in\\mathbb{R}$). That is a field $\\phi:Q\\to \\mathcal{V}$ with a discrete manifold $Q=L$ and value space $\\mathcal{V}=\\mathbb{R}$. Thus, taking $\\phi:Q\\to \\mathcal{V}$ seriously as a fundamental field leads us to thinking of $Q=L$ as the theory's underlying manifold and $\\mathcal{V}=\\mathbb{R}$ as the theory's value space. It is important to note that here, $Q$ is the entire manifold, it is not being thought of as embedded in some larger manifold. (However, a view like this will be considered in Sec.~\\ref{SecExtPart1}.) \n\nTaking $Q$ to be these theories' underlying manifold has consequences for our understanding of the locality of these theories. In a highly intuitive sense, theory KG1 is the most local in that it couples together the fewest lattice sites (only nearest neighbors). Following this KG2 is the next most local in the same sense: it couples only next-to-nearest neighbors. Finally, in this sense KG3 is the least local, it has an infinite range coupling. As mentioned above, there is some tension however with the rate we expect each of these theories to converge at in the continuum limit. How is it that our most non-local theories are somehow the nearest to our perfectly local continuum theory? This first interpretation can do little to resolve this tension, I refer the interested reader to \\cite{DiscreteGenCovPart1} for further discussion.\n\n\\subsection{Intuitive Symmetries}\nWith this manifold $Q=L$ and value space $\\mathcal{V}=\\mathbb{R}$ picked out, what can we expect of these theories' symmetries? For any spacetime theory there are roughly three kinds of symmetries: 1) external symmetries associated with automorphisms of the manifold, here $\\text{Auto}(Q)$, 2) internal symmetries associated with automorphisms of the value space, here $\\text{Auto}(\\mathcal{V})$, and gauge symmetries which result from allowing these internal symmetries to vary smoothly across the manifold. But what are the relevant notions of automorphism here?\n\nAnswering this question for $\\text{Auto}(Q)$ will require us to distinguish what structures are ``built into'' $Q$ and what are ``built on top of'' $Q$. The analogous distinction in the continuum case is that we generally take the manifold's differentiable structure to be built into it while the Minkowski metric, for instance, is something additional built on top of the manifold. In this paper, I am officially agnostic on where we draw this line in the discrete case. However, for didactic purposes I will here be as conservative as possible giving $Q$ as little structure as is sensible. Note that the less structure we associate with $Q$ the larger the class of relevant automorphisms $\\text{Auto}(Q)$ will be. Thus, I am taking $\\text{Auto}(Q)$ to be as large as it can reasonably be. \n\nHere the minimal structure we can reasonably associate with $Q=L$ is that of a set. As such the largest $\\text{Auto}(Q)$ could reasonably be is permutations of the lattice sites, \\mbox{$\\text{Auto}(Q)=\\text{Perm}(L)$}. \n\nIn addition to $\\text{Auto}(Q)$ we might also have internal symmetries $\\text{Auto}(\\mathcal{V})$ and gauge symmetries. While in general there may be abundant internal or gauge symmetries, for the present cases there are not many. In particular, for all of the above-mentioned theories we only have $\\mathcal{V}=\\mathbb{R}$. As mentioned following Eq.~\\eqref{PhiDef}, our (potentially off-shell) discrete fields are themselves vectors $\\phi_\\ell\\in F_L$. Namely, they are closed under addition and scalar multiplication and hence form a vector space. This addition and scalar multiplication is carried out lattice-site-by-lattice-site. Thus, the field's value space $\\mathcal{V}=\\mathbb{R}$ is also structured like a vector space. \n\nThe value space $\\mathcal{V}=\\mathbb{R}$ may additionally have more structure than this. However, as above, for didactic purposes I will here minimize the assumed structure in order to maximize possible symmetries. We can even drop the zero vector from our consideration taking $\\mathcal{V}=\\mathbb{R}$ to be an affine vector space.\nTherefore, I will take \\mbox{$\\text{Auto}(\\mathcal{V})=\\text{Aff}(\\mathbb{R})$} such that our internal symmetries are linear-affine rescalings of $\\phi_\\ell$, namely $\\phi_\\ell\\mapsto c_1\\phi_\\ell+c_2$. We can find the theory's gauge symmetries by letting \\mbox{$c_1,c_2\\in\\mathbb{R}$} vary smoothly across $Q$. That is, $\\phi_\\ell\\mapsto c_{\\ell,1}\\phi_\\ell+c_{\\ell,2}$.\n\nThus, in total, for KG1-KG7 the widest scope of symmetry transformations available to us (at least on this interpretation) are:\n\\begin{align}\\label{PermutationLong}\ns:\\quad \\phi_\\ell\\mapsto c_{P(\\ell),1} \\phi_{P(\\ell)}+c_{P(\\ell),2}\n\\end{align}\nfor some permutation $P:L\\to L$.\n\nFor later reference it will be convenient to translate these potential symmetry transformations in terms of the vector, $\\bm{\\Phi}\\in\\mathbb{R}^L$, as \n\\begin{align}\\label{Permutation}\ns:\\quad\\bm{\\Phi}\\mapsto C_1\\,P\\,\\bm{\\Phi}+\\bm{c}_2,\n\\end{align}\nfor some permutation matrix, $P$, a diagonal matrix $C_1$ and a vector $\\bm{c}_2$. Here $P$ captures the theory's possible external symmetries: the possibility of permuting lattice sites. The diagonal matrix $C_1$ and the vector $\\bm{c}_2$ capture the theory's possible gauge symmetries: the possibility of linear-affine rescalings of $\\phi_\\ell$ which vary smoothly across $Q$.\n\nI will next discuss which transformations of this form preserve the dynamics of KG1-KG7. It should be clear from the outset however that (at least on this interpretation) these theories cannot have continuous spacial translation and rotation, let alone Lorentzian boost symmetries. Indeed, I have been charitable considering the lattice sites structured only as a set (perhaps artificially) increasing the size of $\\text{Auto}(Q)$. Given this, it would be highly surprising if we found KG1-KG7 to have symmetries outside of this set. (Such a surprise is coming in the Sec.~\\ref{SecKlein2}.) \n\nAs I will show in Sec.~\\ref{SecKlein2}, this first interpretation of these theories systematically under predicts the symmetries that discrete spacetime theories can and do have. Fixing this issue will lead one to a discrete analog of general covariance. We here under-predict symmetries because we are taking these theories' lattice structures too seriously. Properly understood, they are merely a coordinate-like representational artifact and so do not limit our symmetries. Before that however, let's see the symmetries these theories have on this interpretation.\n\n\\subsubsection*{Symmetries of KG1-KG7: First Attempt}\nWhat then are the symmetries of KG1-KG7 according to this interpretation? A technical investigation of the symmetries of KG1-KG7 on this interpretation is carried out in Appendix~\\ref{AppA}, but the results are the following. For KG1-KG3 the dynamical symmetries of the form Eq.~\\eqref{PermutationLong} are:\n\\begin{flushleft}\\begin{enumerate}\n \\item[1)] discrete shifts which map lattice site \\mbox{$(j,n)\\mapsto (j-d_1,n-d_2)$} for some integers $d_1,d_2\\in\\mathbb{Z}$,\n \\item[2)] two negation symmetries which map lattice site $(j,n)\\mapsto (-j,n)$ and $(j,n)\\mapsto (j,-n)$ respectively,\n \\item[3)] global linear rescaling which maps \\mbox{$\\phi_\\ell\\mapsto c_1\\phi_\\ell$} for some $c_1\\in\\mathbb{R}$,\n \\item[4)] local affine rescaling which maps \\mbox{$\\phi_\\ell\\mapsto \\phi_\\ell + c_{2,\\ell}(t)$} for some $c_{2,\\ell}(t)$ which is also a solution of the dynamics.\n\\end{enumerate}\\end{flushleft}\nThese are the symmetries of a uniform 1D lattices in space and a uniform 1D lattice in time, \\mbox{$z_{j,n}=(j,n)\\in\\mathbb{R}^2$} (plus linear-affine rescalings). These are two independent 1D lattices (rather than a single square 2D lattice) because we do not have quarter rotations between space and time among our symmetries. Previously I had warned against prematurely interpreting the lattice sites underlying KG1-KG3 as being organized into a square lattice. As it turns out, having investigated these theories' dynamical symmetries this warning was warranted.\n\nWhat about KG4 and KG6? For KG4 and KG6 the dynamical symmetries of the form Eq.~\\eqref{PermutationLong} are:\n\\begin{flushleft}\\begin{enumerate}\n \\item[1)] discrete shifts which map lattice site \\mbox{$(j,n,m)\\mapsto (j-d_1,n-d_2,m-d_3)$} for some integers $d_1,d_2,d_3\\in\\mathbb{Z}$,\n \\item[2)] three negation symmetries which map lattice site $(j,n,m)\\mapsto (-j,n,m)$ and $(j,n,m)\\mapsto (j,-n,m)$ and $(j,n,m)\\mapsto (j,n,-m)$ respectively,\n \\item[3)] a 4-fold symmetry which maps lattice site \\mbox{$(j,n,m)\\mapsto (j,m,-n)$}, \n \\item[4)] global linear rescaling which maps \\mbox{$\\phi_\\ell\\mapsto c_1\\phi_\\ell$} for some $c_1\\in\\mathbb{R}$,\n \\item[5)] local affine rescaling which maps \\mbox{$\\phi_\\ell\\mapsto \\phi_\\ell + c_{2,\\ell}(t)$} for some $c_{2,\\ell}(t)$ which is also a solution of the dynamics.\n\\end{enumerate}\\end{flushleft}\nThese are the symmetries of a square 2D lattice in space and a uniform 1D lattice in time, \\mbox{$z_{j,n,m}=(j,n,m)\\in\\mathbb{R}^3$} (plus linear-affine rescalings). The above 4-fold symmetry corresponds to quarter rotation in space. These are two independent lattices (rather than a single cubic 3D lattice) because we do not have quarter rotations between space and time among our symmetries. Previously I had warned against prematurely interpreting the lattice sites underlying KG4 and KG6 as being organized into a cubic lattice. As it turns out, having investigated these theories' dynamical symmetries this warning was warranted.\n\nWhat about KG5 and KG7? For KG5 and KG7 the dynamical symmetries of the form Eq.~\\eqref{PermutationLong} are:\n\\begin{flushleft}\\begin{enumerate}\n \\item[1)] discrete shifts which map lattice site \\mbox{$(j,n,m)\\mapsto (j-d_1,n-d_2,m-d_3)$} for some integers $d_1,d_2,d_3\\in\\mathbb{Z}$,\n \\item[2)] an exchange symmetry which maps lattice site $(j,n,m)\\mapsto (j,m,n)$,\n \\item[3)] a 6-fold symmetry which maps lattice site \\mbox{$(j,n,m)\\mapsto (j,-m,n+m)$}. (Roughly, this permutes the three terms in the right hand side of Eq.~\\eqref{DKG5} for KG5 and Eq.~\\eqref{DKG7} for KG7.),\n \\item[4)] global linear rescaling which maps \\mbox{$\\phi_\\ell\\mapsto c_1\\phi_\\ell$} for some $c_1\\in\\mathbb{R}$,\n \\item[5)] local affine rescaling which maps \\mbox{$\\phi_\\ell\\mapsto \\phi_\\ell + c_{2,\\ell}(t)$} for some $c_{2,\\ell}(t)$ which is also a solution of the dynamics.\n\\end{enumerate}\\end{flushleft}\nThese are the symmetries of a hexagonal 2D lattice in space and a uniform 1D lattice in time, \\mbox{$z_{j,n,m}=(j,n+m\/2,\\sqrt{3}m\/2)\\in\\mathbb{R}^3$} (plus linear-affine rescalings). The above 6-fold symmetry corresponds to one-sixth rotation in space. Previously I had warned against prematurely interpreting the lattice sites underlying KG5 and KG7 as being organized into a cubic 3D lattice, prompted by the convenient relabeling $\\ell\\mapsto (j,n,m)$. As it turns out, having investigated these theories' dynamical symmetries this warning was well warranted.\n\nThus, by investigating these theories' dynamical symmetries we were able to find what sort of lattice structure the assumed-to-be unstructured lattice $L$ actually has for each theory (e.g. in space a uniform 1D lattice, a square lattice, and a hexagonal lattice each together with a uniform 1D lattice in time).\n\n\\vspace{0.25cm}\n\nFinally, in this interpretation what sense can be made of KG6 and KG7 having a nice one-to-one correspondence between their solutions discussed at the end of Sec.~\\ref{SecSevenKG}? While this correspondence between solutions certainly exists, little sense can be made of it here in support of the equivalence of these theories. As the above discussion has revealed this interpretation associates very different symmetries to KG6 and KG7 and correspondingly very different lattice structures. While there is nothing technically wrong per se with this assessment our later interpretations will make better sense of this correspondence.\n\n\\vspace{0.25cm}\n\nTo summarize, this interpretation has the benefit of being highly intuitive. Taking the fields given to us, \\mbox{$\\phi:Q\\to\\mathbb{R}$}, seriously we identified the underlying manifold as $Q=L$. From this we got some intuitive notions of locality. Moreover, by finding these theories' dynamical symmetries we were able to grant some more structure to their lattice sites. By and large, the interpretation seems to validate all of the first intuitions laid out in Sec.~\\ref{SecIntro}. On this interpretation, the lattice seems to play a substantive role in the theory: it seems to restrict our symmetries, it seems to distinguish our theories from one another, be essentially ``baked-into'' the formalism. (As I will discuss in the next section, none of this is right.)\n\nHowever, there are three major issues with this interpretation which will become clear in light of our later interpretations. Firstly, our locality assessments are in tension with the rates at which these theories converge to the (perfectly local) continuum theory in the continuum limit, see \\cite{DiscreteGenCovPart1} for further discussion. Secondly, despite the niceness of the one-to-one correspondence between the solutions to KG6 and KG7, this interpretation regards them as significantly different theories: with different lattice structures and (here consequently) different symmetries. The final issue (which will become clear in the next section) is that this interpretation drastically under predicts the kinds of symmetries which KG1-KG7 can and do have. In fact, each of KG1-KG7 have a hidden continuous translation symmetry. Moreover, KG6 and KG7 have a hidden continuous rotation symmetry. Moreover still, KG3, KG6, and KG7 all have a hidden (limited) Lorentzian boost symmetry.\n\nAs I will discuss, the root of all of these issues is taking the theory's lattice structure too seriously. As I will argue, when properly understood, they are merely a coordinate-like representational artifact. Indeed, as I will show in the next section they do not limit our theory's symmetries. Moreover, theories appearing initially with different lattice structures may nonetheless be equivalent. Finally, these theories can always be reformulated to refer to no lattice structure at all. These three points establish a strong analogy between the lattice structures appearing in our discrete spacetime theories and the coordinate systems appearing in our continuum theories. Ultimately, spelling out this analogy in detail in Sec.~\\ref{SecDisGenCov} will give us a discrete analog of general covariance (now extended to a Lorentzian context). Indeed, this will lead us to a perfectly Lorentzian lattice theory in Sec.~\\ref{PerfectLorentz}.\n\n\\section{A Second Attempt at Interpreting KG1-KG7}\\label{SecKlein2}\nIn the previous section, I claimed that KG1-KG7 have hidden continuous translation and rotation symmetries and even (limited) Lorentzian boost symmetries. But how can this be? How can discrete spacetime theories have such continuous symmetries? As I discussed in the previous section, if we take our underlying manifold to be \\mbox{$Q=L$} then these theories clearly cannot support continuous translation and rotation symmetries let alone Lorentzian boosts.\n\nIn order to avoid this conclusion we must deny the premise, $Q$ must not be the underlying manifold. What led us to believe $Q$ was the underlying manifold? We arrived at this conclusion by focusing on $\\phi_\\ell\\in F_L$ and thereby taking the real scalar field $\\phi:Q\\to\\mathcal{V}$ to be fundamental. $Q$ is the underlying manifold because it is where our fundamental field maps from. In order to avoid this conclusion we must deny the premise, the field $\\phi:Q\\to\\mathcal{V}$ must not be fundamental. \n\nBut if $\\phi:Q\\to\\mathcal{V}$ is not fundamental then what is? Fortunately, our above discussion has already provided us with another object which we might take as fundamental. Namely, $\\bm{\\Phi}$ defined in Eq.~\\eqref{PhiDef}. These vectors $\\bm{\\Phi}\\in\\mathbb{R}^L$ are in a one-to-one correspondence with the discrete fields $\\phi_\\ell\\in F_L$, Moreover, these vector spaces are isomorphic $\\mathbb{R}^L\\cong F_L$ with Eq.~\\eqref{PhiDef} being a vector space isomorphism between them.\n\nOn this second interpretation I will be taking the formulations of KG1-KG7 in terms of $\\bm{\\Phi}$ seriously: namely Eq.~\\eqref{DKG1}, Eq.~\\eqref{DKG2}, Eq.~\\eqref{DKG3}, and Eqs.~\\eqref{DKG4}-\\eqref{DKG7}. Taken literally as written, what are these theories about? These theories are intuitively about an infinite dimensional vector ($\\bm{\\Phi}\\in\\mathbb{R}^L$) which satisfies some dynamical constraint. \n\nThere are two ways in which one might try to make sense of $\\bm{\\Phi}$ as a field $\\bm{\\Phi}:\\mathcal{M}\\to\\mathcal{V}$ with some manifold $\\mathcal{M}$ and some value space $\\mathcal{V}$. Firstly, one might consider a one-point manifold $\\mathcal{M}=p$ with $\\bm{\\Phi}$ simply being the field value there $\\bm{\\Phi}\\coloneqq \\bm{\\Phi}(p)$. Alternatively, one might try to make sense of $\\bm{\\Phi}$ as a field ``from nowhere'' with an empty manifold $\\mathcal{M}=\\emptyset$. On this view $\\bm{\\Phi}$ is a vector field which takes in no input and returns the vector $\\bm{\\Phi}$.\n\nIn any case, on this interpretation $Q$ is no longer the underlying manifold for KG1-KG7. Indeed, on this interpretation the lattice sites, $L$, are no longer our spacetime manifold (there may not even be a manifold). Rather, they have been ``internalized'' into the value space $\\mathcal{V}=\\mathbb{R}^L$. In particular, in defining this vector space we have associated with each lattice site $\\ell\\in L$ a basis vector $\\bm{b}_\\ell$. See Eq.~\\eqref{PhiDef}. However, as I will discuss, these particular basis vectors play no special role in these theories. Indeed, looking back at the dynamics for each of KG1-KG7 written in terms of $\\bm{\\Phi}$, one can see that in each case it can be made basis-independent.\n\nLet's see what consequences this interpretive stance has for these theories' locality and symmetry. To preview: this second interpretation either dissolves or resolves all of our issues with the first interpretation. Firstly, the tension is dissolved between our theories' differences in locality and their differences in convergence rate in the continuum limit. With no spacetime manifold we no longer have access to any spaciotemporal notion of locality. There simply are no differences in locality anymore. I refer the interested reader to \\cite{DiscreteGenCovPart1} for further discussion.\n\nSecondly, KG6 and KG7 are seen to be equivalent in a stronger sense. And thirdly, perhaps most importantly, this interpretation reveals KG1-KG7's hidden continuous translation and rotation symmetries and even (limited) Lorentzian boost symmetries. However, as I will discuss, this interpretation has some issues of its own which will ultimately require us to make a third attempt at interpreting these theories in Sec.~\\ref{SecKlein3}.\n\n\\subsection{Internalized Symmetries}\nHow does this internalization move affect our theory's capacity for symmetry? How can we now have continuous translation and rotation symmetries as well as a Lorentzian boost symmetry? At first glance, this may appear to have made things worse. Without a manifold (or even if the manifold is a single point) we no longer have any possibility for external symmetries. However, while there are certainly less possible external symmetries, we are now open to a wider range of internal symmetries. It is among these internal symmetries that we will find KG1-KG7's hidden continuous translation and rotation symmetries and even (limited) Lorentzian boost symmetries. As I will argue these symmetries can reasonably be given these names despite being internal symmetries. (In Sec.~\\ref{SecKlein3} I will present a third attempt at interpreting these theories which ``externalizes'' these symmetries, making them genuinely spacial translations, rotations, and Lorentzian boosts.)\n\nWith our focus now on $\\bm{\\Phi}\\in\\mathbb{R}^L$, let us consider its possibilities for symmetries. As discussed above, we have no possibility of external symmetries associated with the manifold. However, we do have possible internal symmetries associated with the value space (i.e., an infinite dimensional vector space) we now have the full range of invertible linear-affine transformations over $\\mathbb{R}^L$, namely $\\text{Auto}(\\mathcal{V})=\\text{Aff}(\\mathbb{R}^L)$. There are no gauge symmetries here as there is no longer any manifold for them to smoothly vary across. Thus, in total the possibly symmetries for our theories under this interpretation are,\n\\begin{align}\\label{GaugeVR}\ns:\\quad\\bm{\\Phi}\\mapsto \\Lambda\\,\\bm{\\Phi}+\\bm{c}\n\\end{align}\nfor some invertible linear transformation $\\Lambda\\in\\text{GL}(\\mathbb{R}^L)$ and some vector $\\bm{c}\\in \\mathbb{R}^L$. \n\nContrast this with the symmetries available to us on our first interpretation, namely Eq.~\\eqref{Permutation}. We can role this new class of possible symmetries back onto our first interpretation as follows: Note that because $\\mathbb{R}^L \\cong F_L$ we have $\\text{Aff}(\\mathbb{R}^L)\\cong \\text{Aff}(F_L)$. The set of transformations $\\text{Aff}(F_L)$ acting on $\\phi_\\ell\\in F_L$ is much larger than what we previously considered: namely, \\mbox{$\\text{Auto}(\\mathbb{V})=\\text{Aff}(\\mathbb{R})$} varying smoothly over \\mbox{$\\text{Auto}(Q)=\\text{Perm}(L)$}. Indeed, the present interpretation has a significantly wider class of symmetries than before. \n\nMoving back to our second interpretation, our previous class of transformations (i.e, \\mbox{$\\text{Auto}(\\mathbb{V})=\\text{Aff}(\\mathbb{R})$} varying smoothly over \\mbox{$\\text{Auto}(Q)=\\text{Perm}(L)$}) corresponds to only a subset of our present consideration: $\\text{Aff}(\\mathbb{R}^L)$. The difference is that before we could only apply a permutation matrix $P$ followed by a diagonal matrix $C_1$ whereas now we are allowed a general linear transformation $\\Lambda$. \n\nNote that permutation and diagonal matrices are basis-dependent notions. Our first interpretation took the lattice sites $\\ell\\in L$ seriously as a part of the manifold $Q=L$ and this is reflected in its conception of symmetries. Converted into $\\mathbb{R}^L$ this first conception of these theories' possible symmetries gives special treatment to the basis associated with the lattice sites, namely $\\{\\bm{b}_\\ell\\}_{\\ell\\in L}$. In particular, on our first interpretation, our possible symmetries are those of the form Eq.~\\eqref{GaugeVR} which \\textit{additionally} preserve this basis (up to rescaling, and reordering).\n\nThis basis receives no special treatment on this second interpretation. While it is true that $\\{\\bm{b}_\\ell\\}_{\\ell\\in L}$ were used in the initial construction of $\\bm{\\Phi}$, after this they no longer play any special role. We are always free to redescribe $\\bm{\\Phi}$ in a different basis if we wish. Indeed, here any change of basis transformation is of the form Eq.~\\eqref{GaugeVR} and hence a symmetry.\n\nWith this attachment to the basis $\\{\\bm{b}_\\ell\\}_{\\ell\\in L}$ dropped, we now have a wider class of symmetries. Indeed, everything which was previously considered a symmetry will be here as well and possibly more. Perhaps among this more general class of symmetries we may find new continuous symmetries. Let's see.\n\n\\subsubsection*{Symmetries of KG1-KG7: Second Attempt}\nWhich of the above transformations are symmetries for KG1-KG7? A non-exhaustive investigation of the symmetries of KG1-KG7 on this interpretation is carried out in Appendix~\\ref{AppA}, but the results are the following. For KG1 and KG2 the dynamical symmetries of the form Eq.~\\eqref{GaugeVR} include:\n\\begin{flushleft}\\begin{enumerate}\n \\item[1)] action by $T^\\epsilon_\\text{j}$ sending $\\bm{\\Phi}\\mapsto T^\\epsilon_\\text{j} \\bm{\\Phi}$ where \\mbox{$T^\\epsilon_\\text{j}=T^\\epsilon\\otimes\\openone$} with $T^\\epsilon$ defined below. Similarly for $T^\\epsilon_\\text{n}=\\openone\\otimes T^\\epsilon$,\n \\item[2)] two negation symmetries which map basis vectors as $\\bm{e}_j\\otimes\\bm{e}_n\\mapsto \\bm{e}_{-j}\\otimes\\bm{e}_n$ and $\\bm{e}_j\\otimes\\bm{e}_n\\mapsto \\bm{e}_j\\otimes\\bm{e}_{-n}$ respectively,\n \\item[3)] a local Fourier rescaling symmetry which maps \\mbox{$\\bm{\\Phi}(\\omega,k)\\mapsto \\tilde{f}(\\omega,k)\\,\\bm{\\Phi}(\\omega,k)$} for some non-zero complex function $\\tilde{f}(\\omega,k)\\in\\mathbb{C}$ with $\\omega,k\\in[-\\pi,\\pi]$,\n \\item[4)] local affine rescaling which maps \\mbox{$\\bm{\\Phi}\\mapsto \\bm{\\Phi} + \\bm{c}_2$} for some $\\bm{c}_2$ which is also a solution of the dynamics.\n\\end{enumerate}\\end{flushleft}\nThese are exactly the same symmetries that we found on the previous interpretation with two differences: Firstly, global rescaling $\\phi_\\ell\\mapsto c_1 \\phi_\\ell$ has been refined to a local Fourier rescaling. Note that the discrete Fourier transform itself is in $\\text{Aff}(\\mathbb{R}^L)$ and so is in the class of potential symmetries considered here.\n\nSecondly, discrete shifts have been replaced with action by \n\\begin{align}\\label{TDef}\nT^\\epsilon\\coloneqq\\text{exp}(-\\epsilon D)\n\\end{align}\nwith $\\epsilon\\in\\mathbb{R}$ acting on each tensor factor. A straight-forward calculation shows that $T^\\epsilon$ acts on the planewave basis \\mbox{$\\bm{\\Phi}(k)\\coloneqq \\sum_{n\\in\\mathbb{Z}}e^{-\\mathrm{i} k n}\\bm{e}_n$} with $k\\in[-\\pi,\\pi]$ as\n\\begin{align}\n\\nonumber\nT^\\epsilon:\\bm{\\Phi}(k)\\mapsto \\exp(\\mathrm{i} k \\epsilon)\\,\\bm{\\Phi}(k).\n\\end{align}\nUsing \\mbox{$\\bm{e}_m\n=\\frac{1}{2\\pi}\\int_{-\\pi}^\\pi e^{\\mathrm{i}\\, k\\, m}\\, \\bm{\\Phi}(k)\\,\\d k$} we can recover how $T^\\epsilon$ acts on the basis $\\bm{e}_m$ as\n\\begin{align}\nT^\\epsilon: \\bm{e}_m \\mapsto \\sum_{b\\in\\mathbb{Z}} S_m(b+\\epsilon) \\, \\bm{e}_b\n\\end{align}\nwhere \n\\begin{align}\\label{SincDef}\nS(y)=\\frac{\\sin(\\pi y)}{\\pi y}, \\quad\\text{and}\\quad\nS_m(y)=S(y-m), \n\\end{align}\nare the normalized and shifted sinc functions. Note that $S_n(m)=\\delta_{nm}$ for integers $n$ and $m$.\n\nAs I will now discuss, $T^\\epsilon$ can be thought of as a continuous translation operator for three reasons despite it being here classified as an internal symmetry. Note that none of these reasons depend on $T^\\epsilon$ being a symmetry of the dynamics.\n\nFirst note that $T^\\epsilon$ is a generalization of the discrete shift operation in the sense that taking $\\epsilon=d_1\\in\\mathbb{Z}$ reduces action by $T^\\epsilon$ to the map $T^{d_1}:\\bm{e}_m\\mapsto \\bm{e}_{m-d_1}$ on basis vectors and relatedly the map $m\\mapsto m-d_1$ on lattice sites. \n\nSecond note that $T^\\epsilon$ is additive in the sense that $T^{\\epsilon_1}\\,T^{\\epsilon_2}=T^{\\epsilon_1+\\epsilon_2}$. In the language or representation theory $T^\\epsilon$ is a representation of the translation group on the vector space $\\mathbb{R}^\\mathbb{Z}\\cong\\mathbb{R}^L$. In particular, this means \\mbox{$T^{1\/2}\\,T^{1\/2}=T^1$}: there is something we can do twice to move one space forward. The same is true for all fractions adding to one. \n\nThird, recall from the discussion following Eq.~\\eqref{LambdaD} that $D$ is closely related to the continuum derivative operator $\\partial_x$, exactly matching its spectrum for $k\\in[-\\pi,\\pi]$. Recall also that the derivative is the generator of translation, i.e. $h(x-\\epsilon)=\\text{exp}(-\\epsilon\\, \\partial_x) h(x)$. In this sense also \\mbox{$T^\\epsilon=\\text{exp}(-\\epsilon D)$} is a translation operator. More will be said about $T^\\epsilon$ in Sec.~\\ref{SecSamplingTheory}.\n\nThus we have our first big lesson: despite the fact that KG1-KG2 can be represented on a lattice, they nonetheless have a continuous translation symmetry. This continuous translation symmetry was hidden from us on our first interpretation because we there took the lattice to be hard-wired in as a part of the manifold. Here, we do not take the lattice structure so seriously. We have internalized it into the value space where it then disappears from view as just one basis among many.\n\nBefore KG3 had the same symmetries as KG1 and KG2, now it does not. However, allow me to skip over KG3 temporarily. In the previous interpretation the symmetries of KG4 and KG6 matched each other, both being associated with a 2D square lattice. Moreover, the symmetries of KG5 and KG7 matched each other, both being associated with a hexagonal 2D lattice. Here however, these pairings are broken up and a new matching pair is formed between KG6 and KG7. More will be said about this momentarily.\n\nLet's consider KG4 first. For KG4 the dynamical symmetries of the form Eq.~\\eqref{GaugeVR} include:\n\\begin{flushleft}\\begin{enumerate}\n \\item[1)] action by $T^\\epsilon_\\text{j}$ sending $\\bm{\\Phi}\\mapsto T^\\epsilon_\\text{j} \\bm{\\Phi}$. Similarly for $T^\\epsilon_\\text{n}$ and $T^\\epsilon_\\text{m}$,\n \\item[2)] three negation symmetries which map basis vectors as \\mbox{$\\bm{e}_{j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}\\mapsto \\bm{e}_{-j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}$}, and \\mbox{$\\bm{e}_{j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}\\mapsto \\bm{e}_{j}\\otimes\\bm{e}_{-n}\\otimes\\bm{e}_{m}$}, and \\mbox{$\\bm{e}_{j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}\\mapsto \\bm{e}_{j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{-m}$} respectively,\n \\item[3)] a 4-fold symmetry which maps basis vectors as \\mbox{$\\bm{e}_{j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}\\mapsto \\bm{e}_{j}\\otimes\\bm{e}_{m}\\otimes\\bm{e}_{-n}$},\n \\item[4)] a local Fourier rescaling symmetry which maps \\mbox{$\\bm{\\Phi}(\\omega,k_1,k_2)\\mapsto \\tilde{f}(\\omega,k_1,k_2)\\,\\bm{\\Phi}(\\omega,k_1,k_2)$} for some non-zero complex function $\\tilde{f}(\\omega,k_1,k_2)\\in\\mathbb{C}$ with $\\omega,k_1,k_2\\in[-\\pi,\\pi]$,\n \\item[5)] local affine rescaling which maps \\mbox{$\\bm{\\Phi}\\mapsto \\bm{\\Phi} + \\bm{c}_2$} for some $\\bm{c}_2$ which is also a solution of the dynamics.\n\\end{enumerate}\\end{flushleft}\nThese are exactly the symmetries which we found on our first interpretation (plus local Fourier rescaling) but with action by $T^\\epsilon_\\text{j}$, $T^\\epsilon_\\text{n}$ and $T^\\epsilon_\\text{m}$ replacing the discrete shifts. The same discussion following Eq.~\\eqref{TDef} applies here, justifying us calling these continuous translation operations. Thus, KG4 has three continuous translation symmetries despite being initially represented on a lattice, Eq.~\\eqref{KG4Long}.\n\nLet's next consider KG5. For KG5 the dynamical symmetries of the form Eq.~\\eqref{GaugeVR} include:\n\\begin{flushleft}\\begin{enumerate}\n \\item[1)] action by $T^\\epsilon_\\text{j}$ sending $\\bm{\\Phi}\\mapsto T^\\epsilon_\\text{j} \\bm{\\Phi}$. Similarly for $T^\\epsilon_\\text{n}$ and $T^\\epsilon_\\text{m}$,\n \\item[2)] a negation symmetry which maps basis vectors as \\mbox{$\\bm{e}_{j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}\\mapsto \\bm{e}_{-j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}$} and an exchange symmetry which maps basis vectors as \\mbox{$\\bm{e}_{j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}\\mapsto \\bm{e}_{j}\\otimes\\bm{e}_{m}\\otimes\\bm{e}_{n}$},\n \\item[3)] a 6-fold symmetry which maps basis vectors as \\mbox{$\\bm{e}_{j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}\\mapsto \\bm{e}_{j}\\otimes\\bm{e}_{-m}\\otimes\\bm{e}_{n+m}$}. (Roughly, this permutes the three terms in Eq.~\\eqref{DKG5}),\n \\item[4)] a local Fourier rescaling symmetry which maps \\mbox{$\\bm{\\Phi}(\\omega,k_1,k_2)\\mapsto \\tilde{f}(\\omega,k_1,k_2)\\,\\bm{\\Phi}(\\omega,k_1,k_2)$} for some non-zero complex function $\\tilde{f}(\\omega,k_1,k_2)\\in\\mathbb{C}$ with $t\\in\\mathbb{R}$ and $\\omega,k_1,k_2\\in[-\\pi,\\pi]$,\n \\item[5)] local affine rescaling which maps \\mbox{$\\bm{\\Phi}\\mapsto \\bm{\\Phi} + \\bm{c}_2$} for some $\\bm{c}_2$ which is also a solution of the dynamics.\n\\end{enumerate}\\end{flushleft}\nThese are exactly the symmetries which we found on our first interpretation (plus local Fourier rescaling) but with action by $T^\\epsilon_\\text{j}$, $T^\\epsilon_\\text{n}$ and $T^\\epsilon_\\text{m}$ replacing the discrete shifts. The same discussion following Eq.~\\eqref{TDef} applies here, justifying us calling these continuous translation operations. Thus, KG5 has three continuous translation symmetries despite being initially represented on a lattice, Eq.~\\eqref{KG5Long}.\n\nBefore moving on to analyze the symmetries of KG6 and KG7, let's first see what this interpretation has to say about them being equivalent to one another. As noted at the end of Sec.~\\ref{SecSevenKG}, KG6 and KG7 have a nice one-to-one correspondence between their solutions. Allow me to spell this out in detail now.\n\nBefore this, however, it is worth briefly noting a rather weak sense in which each of KG4-KG7 are equivalent to each other. As noted following Eq.~\\eqref{SkewKG7KG6} and Eq.~\\eqref{SkewKG6KG7} there is an approximate one-to-one correspondence between each of these theories in the $\\vert \\omega\\vert,\\vert k_1\\vert,\\vert k_2\\vert\\ll\\pi$ regime as they approach their common continuum limit, KG0. By contrast, as I will show, KG6 and KG7 have an \\textit{exact} one-to-one correspondence over \\textit{the whole of} $\\sqrt{k_1^2+k_2^2}<\\pi$ and indeed more. This includes all of their solutions but not all of $\\omega,k_1,k_2\\in[-\\pi,\\pi]$.\n\nThis one-to-one correspondence is mediated by the transformations Eq.~\\eqref{SkewKG7KG6} and Eq.~\\eqref{SkewKG6KG7}. Let's first rewrite these in terms of \\mbox{$\\bm{\\Phi}\\in\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$} as follows. Consider first the transformation which maps the dispersion relation for KG7 onto the one for KG6 (namely, Eq.~\\eqref{SkewKG7KG6}). Consider its action on the planewave basis \n$\\bm{\\Phi}(\\omega,k_1,k_2)$ with \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$}, namely\n\\begin{align}\n\\nonumber\n\\Lambda_{\\text{KG7}\\to\\text{KG6}}:\\bm{\\Phi}(\\omega,k_1,k_2)\\mapsto \\bm{\\Phi}\\left(\\omega,k_1,\\frac{1}{2}k_1+\\frac{\\sqrt{3}}{2}k_2\\right) \n\\end{align}\nA straight-forward calculation shows this acts on the \\mbox{$\\bm{e}_j\\otimes\\bm{e}_n\\otimes\\bm{e}_m$} basis as:\n\\begin{align}\\label{SkewKG7KG6Basis}\n&\\Lambda_{\\text{KG7}\\to\\text{KG6}}:\\bm{e}_j\\otimes\\bm{e}_n\\otimes\\bm{e}_m\\mapsto \\\\\n\\nonumber\n&\\bm{e}_j\\otimes\\left(\\sum_{b_1,b_2\\in\\mathbb{Z}}S_n(b_1+b_2\/2) \\,S_m(\\sqrt{3}\\, b_2\/2)\\, \\bm{e}_{b_1}\\otimes\\bm{e}_{b_2}\\right). \n\\end{align}\nConsider also the transformation which maps the dispersion relation for KG6 onto the one for KG7 (namely, Eq.~\\eqref{SkewKG6KG7}). Consider its action on the planewave basis \n$\\bm{\\Phi}(\\omega,k_1,k_2)$ with \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$}, namely as\n\\begin{align}\n\\nonumber\n\\Lambda_{\\text{KG6}\\to\\text{KG7}}:\\bm{\\Phi}(\\omega,k_1,k_2)\\mapsto \\bm{\\Phi}\\left(\\omega,k_1,\\frac{2k_2-k_1}{\\sqrt{3}}\\right) \n\\end{align}\nA straight-forward calculation shows this acts on the \\mbox{$\\bm{e}_j\\otimes\\bm{e}_n\\otimes\\bm{e}_m$} basis as:\n\\begin{align}\n&\\Lambda_{\\text{KG6}\\to\\text{KG7}}:\\bm{e}_j\\otimes\\bm{e}_n\\otimes\\bm{e}_m\\mapsto \\\\\n\\nonumber\n&\\bm{e}_j\\otimes\\left(\\sum_{b_1,b_2\\in\\mathbb{Z}}S_n(b_1-b_2\/\\sqrt{3}) \\,S_m(2\\, b_2\/\\sqrt{3})\\, \\bm{e}_{b_1}\\otimes\\bm{e}_{b_2}\\right). \n\\end{align}\n\nIt should be noted however that despite the fact that Eq.~\\eqref{SkewKG7KG6} and Eq.~\\eqref{SkewKG6KG7} are each other's inverses, $\\Lambda_{\\text{KG7}\\to\\text{KG6}}$ and $\\Lambda_{\\text{KG6}\\to\\text{KG7}}$ are not each other's inverses (at least not on the whole of \\mbox{$\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$}). This is due to the $2\\pi$ periodic identification of the planewaves $\\bm{\\Phi}(\\omega,k_1,k_2)$. Indeed, when viewed as acting on $\\mathcal{V}=\\mathbb{R}^L$, the transformation $\\Lambda_{\\text{KG7}\\to\\text{KG6}}$ is not even invertible. They are only each other's inverses when we have \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$} both before and after these transformations.\n\nFor these reasons we need to consider the following two subspaces of \\mbox{$\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$}: \n\\begin{align}\n\\mathbb{R}^L_\\text{KG7}\\!\\coloneqq\\text{span}(&\\bm{\\Phi}(\\omega,k_1,k_2)\\vert\\text{with }\\omega,k_1,k_2\\in[-\\pi,\\pi]\\\\\n\\nonumber\n&\\text{ before and after applying Eq.~\\eqref{SkewKG7KG6}})\\\\\n\\mathbb{R}^L_\\text{KG6}\\!\\coloneqq\\text{span}(&\\bm{\\Phi}(\\omega,k_1,k_2)\\vert\\text{with }\\omega,k_1,k_2\\in[-\\pi,\\pi]\\\\\n\\nonumber\n&\\text{ before and after applying Eq.~\\eqref{SkewKG6KG7}}).\n\\end{align}\nFor later reference it should be noted that the rotation invariant subspace,\n\\begin{align}\\label{RLrotinv}\n\\mathbb{R}^L_\\text{rot.inv}\\coloneqq\\text{span}\\left(\\bm{\\Phi}(\\omega,k_1,k_2)\\Big\\vert\\,\\sqrt{k_1^2+k_2^2}<\\pi\\right)\n\\end{align}\nis a subspace of $\\mathbb{R}^L_\\text{KG6}$, that is $\\mathbb{R}^L_\\text{rot.inv}\\subset \\mathbb{R}^L_\\text{KG6}$.\n\nRestricted to $\\mathbb{R}^L_\\text{KG6}$ and $\\mathbb{R}^L_\\text{KG7}$ these transformations are invertible and indeed are each other's inverses. Fortunately, all of KG6's solutions are in $\\mathbb{R}^L_\\text{KG6}$ (and moreover they are in $\\mathbb{R}^L_\\text{rot.inv}$ as well). Similarly, all of KG7's solutions are in $\\mathbb{R}^L_\\text{KG7}$. Thus, $\\Lambda_{\\text{KG7}\\to\\text{KG6}}$ maps generic solutions to KG7 onto generic solutions for KG6 in an invertible way. Therefore, $\\Lambda_{\\text{KG6}\\to\\text{KG7}}$ and $\\Lambda_{\\text{KG7}\\to\\text{KG6}}$ give us not only a one-to-one correspondence between the solutions to KG6 and KG7 but a solution-preserving vector-space isomorphism between KG6 and KG7. One can gloss this situation saying: the KPMs of KG6 and KG7 are not isomorphic, but their DPMs are.\n\nThe fact that this is solution-preserving vector-space isomorphism rather than merely a one-to-one correspondence has substantial consequences for these theories' symmetries. Namely, this forces KG6 and KG7 to have the same symmetries. This is because these transformations are both of the form Eq.~\\eqref{GaugeVR} (but notably not of the form Eq.~\\eqref{Permutation}) for any symmetry transformation for KG6 there is a corresponding symmetry transformation for KG7 and vice versa. \n\nThis is in strong contrast to the results of our previous analysis in Sec.~\\ref{SecKlein1}. There KG6 was seen to have symmetries associated with a square 2D lattice and KG7 was seen to have the symmetries associated with a hexagonal 2D lattice. By contrast, in the present interpretation KG6 and KG7 are thoroughly equivalent: We have a solution-preserving vector-space isomorphism between them. Thus, on this interpretation the only difference between KG6 and KG7 is a change of basis.\n\nThus we have our second big lesson: despite the fact that KG6 and KG7 can be represented with very different lattice structures (i.e., a square lattice versus a hexagonal lattice) they have nonetheless turned out to be completely equivalent to one another. This equivalence was hidden from us on our first interpretation because we there took the lattice too seriously. As I will now discuss, this reduced their continuous rotation symmetries down to quarter rotations and one-sixth rotations respectively and thereby made them inequivalent. Here, we do not take the lattice structure so seriously. We have here internalized it into the value space where it subsequently disappears from view as just one basis among many.\n\nIn addition to switching between lattice structures, in this interpretation we can also do away with them altogether. In this interpretation, a lattice structure is associated with a choice of basis for $\\mathcal{V}=\\mathbb{R}^L$. A choice of basis (like a choice of coordinates) is ultimately a merely representational choice, reflecting no physics. We can always choose, if we like, to work within a basis-free formulation of these theories. That is, ultimately, a lattice-free formulation of these theories. Thus we have our third big lesson: given a discrete spacetime theory with some lattice structure we can always reformulate it in such a way that it has no lattice structure whatsoever.\n\nIn the rest of this subsection I will only discuss the symmetries KG6; analogous conclusions are true for KG7 after applying $\\Lambda_{\\text{KG6}\\to\\text{KG7}}$. For KG6 the dynamical symmetries of the form Eq.~\\eqref{GaugeVR} include:\n\\begin{flushleft}\\begin{enumerate}\n \\item[1)] action by $T^\\epsilon_\\text{j}$ sending $\\bm{\\Phi}\\mapsto T^\\epsilon_\\text{j} \\bm{\\Phi}$. Similarly for $T^\\epsilon_\\text{n}$ and $T^\\epsilon_\\text{m}$,\n \\item[2)] three negation symmetries which map basis vectors as \\mbox{$\\bm{e}_{j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}\\mapsto \\bm{e}_{-j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}$}, and \\mbox{$\\bm{e}_{j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}\\mapsto \\bm{e}_{j}\\otimes\\bm{e}_{-n}\\otimes\\bm{e}_{m}$}, and \\mbox{$\\bm{e}_{j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}\\mapsto \\bm{e}_{j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{-m}$} respectively,\n \\item[3)] action by $R^\\theta$ sending $\\bm{\\Phi}\\mapsto R^\\theta \\bm{\\Phi}$ with $R^\\theta$ defined below. (This being a symmetry requires some qualification as I will discuss below.),\n \\item[4)] action by $\\Lambda^w_\\text{j,n}$ sending $\\bm{\\Phi}\\mapsto \\Lambda^w_\\text{j,n} \\bm{\\Phi}$ with $\\Lambda^w_\\text{j,n}$ defined below. Similarly for $\\Lambda^w_\\text{j,m}$ which is defined below as well. (This being a symmetry requires some qualification as I will discuss below.),\n \\item[5)] a local Fourier rescaling symmetry which maps \\mbox{$\\bm{\\Phi}(\\omega,k_1,k_2)\\mapsto \\tilde{f}(\\omega,k_1,k_2)\\,\\bm{\\Phi}(\\omega,k_1,k_2)$} for some non-zero complex function $\\tilde{f}(\\omega,k_1,k_2)\\in\\mathbb{C}$ with $\\omega,k_1,k_2\\in[-\\pi,\\pi]$,\n \\item[6)] local affine rescaling which maps \\mbox{$\\bm{\\Phi}\\mapsto \\bm{\\Phi} + \\bm{c}_2$} for some $\\bm{c}_2$ which is also a solution of the dynamics.\n\\end{enumerate}\\end{flushleft}\nAs with KG4 and KG5, we have here gained local Fourier rescaling and action by $T^\\epsilon_\\text{j}$, $T^\\epsilon_\\text{n}$ and $T^\\epsilon_\\text{m}$ has replaced the discrete shifts from before. The same discussion following Eq.~\\eqref{TDef} applies here, justifying us calling these continuous translation operations. Thus, KG6 (and KG7) have three continuous translation symmetries despite being initially represented on a lattice.\n\nAdditionally, we have the quarter rotation symmetry from our first interpretation replaced with action by $R^\\theta$, which as I will argue is essentially a continuous rotation transformation. Before that, it is worth noting a rather weak sense in which each of KG4-KG7 are rotation invariant. Each of these theories is approximately rotation invariant in the $\\vert \\omega\\vert,\\vert k_1\\vert,\\vert k_2\\vert\\ll\\pi$ regime as they approach the continuum limit. By contrast, as I will show, KG6 is \\textit{exactly} rotation invariant over \\textit{the whole of} $\\sqrt{k_1^2+k_2^2}<\\pi$, that is the whole of $\\mathbb{R}^L_\\text{rot.inv.}$. Since all of KG6's solutions lie inside of $\\mathbb{R}^L_\\text{rot.inv.}$, $R^\\theta$ will always map its solutions to solutions in an invertible way, and will hence be a symmetry.\n\nThis alleged continuous rotation transformation $R^\\theta$ is given by \n\\begin{align}\\label{RthetaDef}\nR^\\theta \\coloneqq \\exp(-\\theta (N_\\text{n} D_\\text{m}-N_\\text{m} D_\\text{n}))\n\\end{align}\nwith $\\theta\\in\\mathbb{R}$ and where $N$ is the ``position operator'' which acts on the basis $\\bm{e}_m$ as $N\\bm{e}_m=m\\,\\bm{e}_m$ for $m\\in\\mathbb{Z}$. Thus \\mbox{$N_\\text{n}=\\openone\\otimes N\\otimes\\openone$} and \\mbox{$N_\\text{m}=\\openone\\otimes \\openone\\otimes N$} return the second and third index respectively when acting on $\\bm{e}_j\\otimes \\bm{e}_n\\otimes \\bm{e}_m$.\n\nA straight-forward calculation shows that $R^\\theta$ acts on the planewave basis $\\bm{\\Phi}(\\omega,k_1,k_2)$ with \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$} as\n\\begin{align}\nR^\\theta:\\,&\\bm{\\Phi}(\\omega,k_1,k_2) \\mapsto\\\\\n\\nonumber\n&\\bm{\\Phi}(\\omega,\\cos(\\theta)k_1-\\sin(\\theta)k_2,\\sin(\\theta)k_1+\\cos(\\theta)k_2).\n\\end{align}\nand acts on the basis $\\bm{e}_j\\otimes \\bm{e}_n\\otimes \\bm{e}_m$ as\n\\begin{align}\n&R^\\theta: \\bm{e}_j\\!\\otimes\\bm{e}_n\\!\\otimes\\bm{e}_m \\mapsto\\bm{e}_{j}\\otimes\\!\\!\\!\\!\\sum_{b_1,b_2\\in\\mathbb{Z}} R_{nm}^{b_1 b_2}(\\theta)\\,\n\\bm{e}_{b_1}\\!\\!\\otimes\\bm{e}_{b_2}\\\\\n\\nonumber\n&R_{nm}^{b_1 b_2}\\!(\\theta)\\!=\\!S_n(\\cos(\\theta) b_1 \\!-\\! \\sin(\\theta) b_2) S_m(\\sin(\\theta) b_1\\!+\\! \\cos(\\theta)b_2).\n\\end{align}\n\nIt should be noted that $R^\\theta$ is not invertible (at least not on the whole of \\mbox{$\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$}). As I will soon discuss, this is due to the $2\\pi$ periodic identification of the planewaves $\\bm{\\Phi}(\\omega,k_1,k_2)$. However, $R^\\theta$ is invertible over $\\mathbb{R}^L_\\text{rot.inv.}$ which contains all of KG6's solutions. Thus, for KG6 $R^\\theta$ always maps solutions to solutions in an invertible way, and is hence a symmetry.\n\nTo see why $R^\\theta$ is not invertible over all of $\\mathcal{V}=\\mathbb{R}^L$ note that $R^{\\pi\/4}$ maps two different planewaves to the same place: Firstly note,\n\\begin{align}\nR^{\\pi\/4}\\bm{\\Phi}(\\omega,\\pi,\\pi)&=\\bm{\\Phi}(\\omega,0,\\sqrt{2}\\pi)\\\\\n\\nonumber\n&=\\bm{\\Phi}(\\omega,0,\\sqrt{2}\\pi-2\\pi)\n\\end{align}\nsince the planewaves repeat themselves with period $2\\pi$. Secondly note, \n\\begin{align}\n\\nonumber\nR^{\\pi\/4}\\bm{\\Phi}(\\omega,\\pi-\\sqrt{2}\\pi,\\pi-\\sqrt{2}\\pi)&=\\bm{\\Phi}(\\omega,0,\\sqrt{2}\\pi-2\\pi).\n\\end{align}\nSuch issues do not arise when $\\sqrt{k_1^2+k_2^2}<\\pi$. Thus, when we restrict our attention to $\\mathbb{R}^L_\\text{rot.inv}$ (which contains all of KG6's solutions) then $R^\\theta$ is invertible and indeed a symmetry. One can gloss this situation saying: rotation does not map the KPMs of KG6 onto themselves in an invertible way, but it does for the DPMs of KG6. If we cut the KPMs of KG6 down to $\\mathbb{R}^L_\\text{rot.inv}$ this minor issue is fixed.\n\nAs I will now discuss, $R^\\theta$ can be thought of as a continuous rotation operator for three reasons despite it being here an internal symmetry. First note that $R^\\theta$ is a generalization of quarter rotation operation in the sense that taking $\\theta=\\pi\/2$ reduces action by $R^\\theta$ to the map \\mbox{$R^{\\pi\/2}:\\bm{e}_{j}\\otimes\\bm{e}_{n}\\otimes\\bm{e}_{m}\\mapsto \\bm{e}_{j}\\otimes\\bm{e}_{m}\\otimes\\bm{e}_{-n}$} on basis vectors and relatedly the map $(j,n,m)\\mapsto (j,m,-n)$ on lattice sites.\n\nSecond, note that restricted to $\\mathbb{R}^L_\\text{rot.inv.}$, $R^\\theta$ is cyclically additive in the sense that $R^{\\theta_1}\\,R^{\\theta_2}=R^{\\theta_1+\\theta_2}$ with $R^{2\\pi}=\\openone$. In the language of representation theory, $R^\\theta$ is a representation of the rotation group on the vector space $\\mathbb{R}^L_\\text{rot.inv.}$. In particular, this means \\mbox{$R^{\\pi\/4}\\,R^{\\pi\/4}=R^{\\pi\/2}$}. There is something we can do twice to make a quarter rotation. Similarly for all fractional rotations. Moreover, note that together with the above discussed translations, these constitute a representation of the Euclidean group over $\\mathbb{R}^L_\\text{rot.inv.}$.\n\nThird, recall from the discussion following Eq.~\\eqref{LambdaD} that $D$ is closely related to the continuum derivative operator $\\partial_x$, exactly matching its spectrum for $k\\in[-\\pi,\\pi]$. Recall also that rotations are generated through the derivative as \\mbox{$h(R^\\theta(x,y))= \\exp(-\\theta (x \\partial_y-y \\partial_x))h(x,y)$}. In this sense also $R^\\theta$ is a rotation operator. More will be said about $R^\\theta$ in Appendix~\\ref{AppA}.\n\nThis adds to our first big lesson: despite the fact that KG6 and KG7 can be represented on a cubic 3D lattice and a hexagonal 3D lattice respectively, they nonetheless both have a continuous rotation symmetry. This, in addition to their continuous translation symmetries. These continuous translation and rotations symmetries were hidden from us on our first interpretation because we there took the lattice representations too seriously. Here, we do not take the lattice structure so seriously. Instead, we have internalized it into the value space where it then disappears from view as just one basis among many.\n\nLet's next discuss KG6's (limited) Lorentzian boost symmetry. In addition to generalizing the 4-fold symmetry into to a continuous rotation symmetry, KG6 also has a brand new symmetry on this interpretation, namely ``action by $\\Lambda^w_\\text{j,n}$ and\/or $\\Lambda^w_\\text{j,m}$''. As I will argue these are essentially Lorentz boost transformations. Before that, it is worth noting a rather weak sense in which each of KG4-KG7 are Lorentz invariant. Each of these theories is approximately Lorentz invariant as they approach the continuum limit regime $\\vert \\omega\\vert,\\vert k_1\\vert,\\vert k_2\\vert\\ll\\pi$ at least boosts parameters $w$ which keep them in this regime. By contrast, as I will show, KG6 is \\textit{exactly} Lorentz invariant over \\textit{a finite-sized region} around $\\omega,k_1,k_2=0$ and boost parameter $w=0$.\n\nThis alleged Lorentz boost transformations are given by\n\\begin{align}\\label{LambdaWDef}\n\\Lambda^w_\\text{j,n}\\coloneqq\\exp(-w (&N_\\text{j} D_\\text{n} + N_\\text{n} D_\\text{j})),\\\\\n\\nonumber\n\\Lambda^w_\\text{j,m}\\coloneqq\\exp(-w (&N_\\text{j} D_\\text{m} + N_\\text{m} D_\\text{j}))\n\\end{align}\nwith $w\\in\\mathbb{R}$. Note that $\\Lambda^w_\\text{j,n}$ acts only on the first and second tensor factor, whereas $\\Lambda^w_\\text{j,m}$ acts only on the first and third factors. In what follows I will focus on $\\Lambda^w_\\text{j,n}$, with similar results following for $\\Lambda^w_\\text{j,m}$.\n\nA straight-forward calculation shows that $\\Lambda^w_\\text{j,n}$ acts on the planewave basis $\\bm{\\Phi}(\\omega,k_1,k_2)$ with \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$} as\n\\begin{align}\n\\Lambda^w_\\text{j,n}:\\,&\\bm{\\Phi}(\\omega,k_1,k_2) \\mapsto\\\\\n\\nonumber\n&\\bm{\\Phi}(\\cosh(w)\\omega+\\sinh(w)k_1,\\sinh(w)\\omega+\\cosh(w)k_1,k_2).\n\\end{align}\nThis is a Lorentz boost in discrete Fourier space. It follows from this that $\\Lambda^w_\\text{j,n}$ acts on the basis $\\bm{e}_j\\otimes \\bm{e}_n\\otimes\\bm{e}_m$ as\n\\begin{align}\n\\nonumber\n&\\Lambda^w_\\text{j,n}: \\bm{e}_j\\otimes\\!\\bm{e}_n\\otimes\\!\\bm{e}_m\\mapsto\\!\\!\\!\\!\\sum_{b_1,b_2\\in\\mathbb{Z}} \\Lambda_{jn}^{b_1 b_2}(w)\\,\n\\bm{e}_{b_1}\\otimes\\!\\bm{e}_{b_2}\\otimes\\!\\bm{e}_m\\\\\n&\\Lambda_{jn}^{b_1 b_2}(w)=S_j(\\cosh(w) b_1 + \\sinh(w) b_2)\\\\ \n\\nonumber\n&\\qquad\\qquad\\times S_n(\\sinh(w) b_1+\\cosh(w) b_2).\n\\end{align}\nLike with $R^\\theta$ discussed above, $\\Lambda^w_\\text{j,n}$ is a symmetries of the dynamics in a qualified sense: namely, $\\Lambda^w_\\text{j,n}$ is not invertible (at least not on the whole of \\mbox{$\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$}). Indeed, each of these transformations is only invertible for a subspace of \\mbox{$\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$}. As before, this is due to the $2\\pi$ periodic identification of the planewaves $\\bm{\\Phi}(\\omega,k_1,k_2)$. To see this note that $\\Lambda^{\\ln(2)}_\\text{j,n}$ maps two different planewaves to the same place: Firstly note,\n\\begin{align}\n\\Lambda^{\\ln(2)}_\\text{j,n}\\bm{\\Phi}(\\pi,\\pi,k_2)&=\\bm{\\Phi}(2\\pi,2\\pi,k_2)\\\\\n\\nonumber\n&=\\bm{\\Phi}(0,0,k_2)\n\\end{align}\nsince the planewaves repeat themselves with period $2\\pi$. Secondly note, \n\\begin{align}\n\\nonumber\n\\Lambda^{\\ln(2)}_\\text{j,n}\\bm{\\Phi}(0,0,k_2)&=\\bm{\\Phi}(0,0,k_2)\n\\end{align}\nSuch issues do not arise when we have \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$} both before and after the Lorentz transformation. \n\nThus, $\\Lambda^{w}_\\text{j,n}$ and $\\Lambda^{w}_\\text{j,m}$ are invertible over the portion of \\mbox{$\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$} spanned by planewaves $\\bm{\\Phi}(\\omega,k_1,k_2)$ which satisfy the following two conditions:\n\\begin{align}\\label{BoostCond}\n\\vert\\cosh(w)\\,\\omega+\\sinh(w)\\, k_1\\,\\vert<&\\pi\\\\\n\\nonumber\n\\vert\\sinh(w)\\,\\omega+\\cosh(w)\\, k_2\\,\\vert<&\\pi.\n\\end{align}\n\nUnfortunately, unlike with $R^\\theta$ this issue cannot be so easily contained: the region where both $\\Lambda^{w}_\\text{j,n}$ and $\\Lambda^{w}_\\text{j,m}$ are invertible depends on $w$ and indeed, shrinks to nothing as $w\\to\\pm\\infty$. When boosted enough, any planewave except $\\omega=k_1=k_2=0$ will leave the region \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$}. Moreover, for any $w$ there is some planewave in \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$} which when boosted by $w$ leaves this region. However, despite this, for every planewave in $\\omega,k_1,k_2\\in(-\\pi,\\pi)$ there is some small enough boost $w\\neq0$ which keeps it in \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$}. Moreover, for every \\mbox{$w$} there is some neighborhood around \\mbox{$\\omega,k_1,k_2=0$} where boosting by $w$ leaves us in \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$}. When restricted to acting on the span of these planewaves $\\Lambda^{w}_\\text{j,n}$ and $\\Lambda^{w}_\\text{j,n}$ are both invertible.\n\nIt is in the following sense that $\\Lambda^{w}_\\text{j,n}$ and $\\Lambda^{w}_\\text{j,m}$ are symmetries of KG6. Consider solutions to KG6 which are supported only over planewaves in some finite neighborhood of $\\omega,k_1,k_2=0$. Consider $\\Lambda^{w}_\\text{j,n}$ and $\\Lambda^{w}_\\text{j,m}$ with $w$ in some finite neighborhood of $w=0$. Choose the neighborhoods such that Eq.~\\eqref{BoostCond} is satisfied throughout. Restricting our attention to this regime, all of these transformations map exactly map solutions to solutions in an invertible way. As I claimed above, KG6 is \\textit{exactly} Lorentz invariant over \\textit{a finite-sized region} around $\\omega,k_1,k_2=0$ and boost parameter $w=0$.\n\nRecall as mentioned above that for all planewaves in $\\omega,k_1,k_2\\in(-\\pi,\\pi)$ there is some small enough boost \\mbox{$w\\neq0$} which keeps it in \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$}. Thus, KG6 has a differential Lorentz boost invariance over all of \\mbox{$\\omega,k_1,k_2\\in(-\\pi,\\pi)$}. Note that the same is true of translations and rotations. Including translations and rotations, KG6 has a differential Poincar\\'e invariance over $\\omega,k_1,k_2\\in(-\\pi,\\pi)$. Concretely, taken all together $\\Lambda^w_\\text{j,n}$, $\\Lambda^w_\\text{j,m}$, $R^\\theta$, $T^\\epsilon_\\text{j}$, $T^\\epsilon_\\text{n}$ and $T^\\epsilon_\\text{m}$ form a representation of the Poincar\\'e algebra over the space spanned by \\mbox{$\\omega,k_1,k_2\\in(-\\pi,\\pi)$}. In particular, this means that for every algebraic fact about the differential Poincar\\'e transformations there is an analogous fact here with the group action being replaced by matrix multiplication. Exponentiating this representation of the Poincar\\'e algebra we recover a finite part of the Poincar\\'e group. Namely, the finite-sized collection of states and transformations satisfying:\n\\begin{align}\\label{PoincareCond}\n\\vert\\cosh(w)\\,\\omega+\\sinh(w)\\,\\vert k\\vert\\,\\vert<&\\pi\\\\\n\\nonumber\n\\vert\\sinh(w)\\,\\omega+\\cosh(w)\\,\\vert k\\vert\\,\\vert<&\\pi\\\\\n\\nonumber\n\\vert\\cos(\\theta)\\,k_1-\\sin(\\theta) k_2\\vert<&\\pi\\\\\n\\nonumber\n\\vert\\sin(\\theta)\\,k_1+\\cos(\\theta) k_2\\vert<&\\pi.\n\\end{align}\nIndeed, KG6 is \\textit{exactly} Poincar\\'e invariant over \\textit{a finite-sized region} around $\\omega,k_1,k_2=0$ and $w=\\theta=0$.\n\nAs I will now discuss, in this regime $\\Lambda^w_\\text{j,m}$ and $\\Lambda^w_\\text{j,n}$ can be thought of as implementing Lorentzian boosts for two reasons despite being here categorized as an internal symmetry. Firstly, recall the close relationship noted above between $D$ and $\\partial_x$. Recall also that Lorentz boosts are generated through the derivative as \\mbox{$h(\\Lambda^w(t,x))= \\exp(-w (x \\partial_t + t \\partial_x))h(t,x)$}. \n\nSecondly, as I have already mentioned, together with our above discussed translations and rotations, $\\Lambda^w_\\text{j,m}$ and $\\Lambda^w_\\text{j,n}$ give us a representation of the Poincar\\'e algebra and even a finite portion of the Poincar\\'e group.\n\nThus we have yet another addendum to our first big lesson: despite the fact that KG6 and KG7 can be represented on a cubic 3D lattice and a hexagonal 3D lattice respectively, they nonetheless both have a Lorentzian boost symmetries (in a finite but limited regime). This, in addition to their continuous translation and rotation symmetries. As impressive as this admittedly limited Lorentz symmetry is, we can do better: In Sec.~\\ref{PerfectLorentz} I will provide a perfectly Lorentzian lattice theory.\n\nFinally, let's consider KG3. For KG3 the dynamical symmetries of the form Eq.~\\eqref{GaugeVR} include:\n\\begin{flushleft}\\begin{enumerate}\n \\item[1)] action by $T^\\epsilon_\\text{j}$ sending $\\bm{\\Phi}\\mapsto T^\\epsilon_\\text{j} \\bm{\\Phi}$. Similarly for $T^\\epsilon_\\text{n}$.\n \\item[2)] two negation symmetries which map basis vectors as \\mbox{$\\bm{e}_{j}\\otimes\\bm{e}_{n}\\mapsto \\bm{e}_{-j}\\otimes\\bm{e}_{n}$} and \\mbox{$\\bm{e}_{j}\\otimes\\bm{e}_{n}\\mapsto \\bm{e}_{j}\\otimes\\bm{e}_{-n}$} respectively,\n \\item[3)] action by $\\Lambda^w_\\text{jn}$ sending $\\bm{\\Phi}\\mapsto \\Lambda^w_\\text{jn} \\bm{\\Phi}$ where \\mbox{$\\Lambda^w_\\text{jn}$} is defined above,\n \\item[4)] a local Fourier rescaling symmetry which maps \\mbox{$\\bm{\\Phi}(\\omega,k)\\mapsto \\tilde{f}(\\omega,k)\\,\\bm{\\Phi}(\\omega,k)$} for some non-zero complex function $\\tilde{f}(\\omega,k)\\in\\mathbb{C}$ with $\\omega,k\\in[-\\pi,\\pi]$,\n \\item[5)] local affine rescaling which maps \\mbox{$\\bm{\\Phi}\\mapsto \\bm{\\Phi} + \\bm{c}_2$} for some $\\bm{c}_2$ which is also a solution of the dynamics.\n\\end{enumerate}\\end{flushleft}\nAs with KG1 and KG2 we have gained local Fourier rescaling and discrete shifts have been replaced with action by $T^\\epsilon$. As discussed above we are justified in calling $T^\\epsilon$ a continuous translation operation. For KG3 we also have a new symmetry, namely action by $\\Lambda^w_\\text{jn}$. For the reasons discussed above, this can be thought of as a (limited) Lorentzian boost symmetry despite it being here classified as an internal symmetry.\n\n\\vspace{0.25cm}\nTo summarize: this second attempt at interpreting KG1-KG7 has fixed all of the issues with our previous interpretation. Firstly, there is no longer any tension between these theories' differing locality properties and the rates at which they converge to the (perfectly local) continuum theory in the continuum limit. (There are no longer any differences in locality.) See \\cite{DiscreteGenCovPart1} for further discussion. Secondly, the fact that we have a nice one-to-one correspondence between the solutions to KG6 and KG7 is now more satisfyingly reflected in their matching symmetries. Finally, this interpretation has exposed the fact that KG1-KG7 have hidden continuous translation and rotation symmetries as well as a (limited) Lorentzian boost symmetries.\n\nBy and large, the interpretation invalidates all of the first intuitions laid out in Sec.~\\ref{SecIntro}. As this interpretation has revealed, the lattice seems to play a merely representational role in the theory: it does not restrict our symmetries. Moreover, theories initially appearing with different lattice structures may nonetheless turn out to be completely equivalent. The process for switching between lattice structures is here a change of basis in the value space. Indeed, we have a third lesson: there is no sense in which these lattice structures are essentially ``baked-into'' these theories; on this interpretation our theories make no reference to any lattice structure if we work in a basis-independent way. No basis is dynamically favored. \n\nAs I discussed in \\cite{DiscreteGenCovPart1} these three lessons lay the foundation for a strong analogy between the lattice structures which appear in our discrete spacetime theories and the coordinate systems which appear in our continuum spacetime theories. This analogy here extended to Lorentzian theories.\n\nThese are substantial lessons, but ultimately this interpretation has a few issues of its own. Firstly, the way that the tension is dissolved between locality and convergence in the continuum limit is unsatisfying. Intuitively, we ought to be able to extract intuitions about locality from the lattice sites. \n\nMoreover, while this interpretation has indeed exposed KG1-KG7's hidden symmetries, the way it classifies them seems wrong. They are here classified as internal symmetries (i.e., symmetries on the value space) whereas intuitively they should be external symmetries (i.e., symmetries on the manifold).\n\nThe root of all of these issues is taking the theory's lattice structure to be internalized into the theory's value space. Our third attempt at interpreting these theories will fix this by externalizing these symmetries. As I will discuss, this gives us access to a perspective within which we can find a perfectly Lorentzian lattice theory (i.e., not limited as the above theories are).\n\n\\section{Externalizing these theories - Part 1}\\label{SecExtPart1}\nIn the previous section it was revealed that KG1-KG7 have hidden continuous symmetry transformations which intuitively correspond to spacial translations and rotation and even (limited) Lorentzian boosts. In our first attempt at interpreting KG1-KG7 the possibility of such symmetries were outright denied, see Sec.~\\ref{SecKlein1}. In our second attempt, these hidden symmetries were exposed, but they were classified (unintuitively) as internal symmetries, see Sec.~\\ref{SecKlein2}. This is due to an ``internalization'' move made in our second interpretation. This move also had the unfortunate consequence of undercutting our ability to use the lattice sites to reason about locality.\n\nIn this section I will show how we can externalize these symmetries by in a principled way 1) inventing a continuous manifold for our formerly discrete theories to live on and 2) embedding our theory's states\/dynamics onto this manifold as a new dynamical field.\n\n\\subsection{A Principled Choice of Spacetime Manifold}\\label{SecKlein3A}\nIf we are going to externalize these symmetries then we need to have a big enough manifold on which to do the job. What spacetime manifold $\\mathcal{M}$ might be up to the task? The first thing we must do is pick out which of our theory's symmetries we would like to externalize (there may be some symmetries we want to keep internal). For KG1-KG7 we want to externalize the following symmetries: continuous translations, mirror reflections, as well as discrete rotations for KG4 and KG5 and continuous rotations for KG6 and KG7. For each theory we can collect these dynamical symmetries together in a group $G^\\text{dym}_\\text{to-be-ext}$. Clearly, our choice of spacetime manifold $\\mathcal{M}$ needs to be big enough to have $G^\\text{dym}_\\text{to-be-ext}$ as a subgroup of $\\text{Diff}(\\mathcal{M})$. Let us call this the symmetry-fitting constraint\\footnote{As I will soon discuss, there may be transformations which we want to externalize even when they are not symmetries.}. \n\nOf course, symmetry-fitting alone doesn't uniquely specify which manifold we ought to use. If $\\mathcal{M}$ works, then so does any larger $\\mathcal{M}'$ with $\\mathcal{M}$ as a sub-manifold. For standard Occamistic reasons, it is natural to go with the smallest manifold which gets the job done. The larger the gap between the groups $G^\\text{dym}_\\text{to-be-ext}$ and $\\text{Diff}(\\mathcal{M})$ the more fixed spacetime structures will need to be introduced later on (see Sec.~\\ref{SecFullGenCov}).\n\nIn principle, we are free to pick any large-enough manifold which we like to embed KG1-KG7 onto. However, perhaps surprisingly, if we make natural choices about how the translation operations we have already identified (see Eq.~\\eqref{TDef}) fit onto the new spacetime manifold then our choice of $\\mathcal{M}$ is actually fixed up to diffeomorphism. In particular, I demand the following: Certain translation operations on $\\mathbb{R}^L$ are to correspond (perhaps in a complicated way) to parallel transport on the new spacetime manifold $\\mathcal{M}$. For KG1-KG3 these to-be-externalized translation operations are $T^{\\epsilon_1}_\\text{j}$ and $T^{\\epsilon_2}_\\text{n}$. For KG4-KG7 these to-be-externalized translation operations are $T^{\\epsilon_1}_\\text{j}$, $T^{\\epsilon_2}_\\text{n}$, and \n$T^{\\epsilon_3}_\\text{m}$. Let us call this the translation-matching constraint. As I will soon show, this constraint fixes the new spacetime manifold $\\mathcal{M}$ up to diffeomorphism.\n\nBefore fleshing this out, however, it's worth reflecting on two questions focusing on KG4-KG7: What exactly makes $T^{\\epsilon_1}_\\text{j}$, $T^{\\epsilon_2}_\\text{n}$, and $T^{\\epsilon_3}_\\text{m}$ translation operations? Moreover, what motivation do we have to externalize these particular translation operations? To answer: firstly, these can be thought of as translation operations for the reasons discussed following Eq.~\\eqref{TDef}. Note these reasons are unrelated to the fact that these are dynamical symmetries of KG4-KG7. Suppose the dynamics of KG4 given by Eq.~\\eqref{KG4Long} was modified to have explicit dependence on the index $n$. In this case, $T^{\\epsilon_2}_\\text{n}$ would no longer be a dynamical symmetry but it would still be a translation operation (and moreover, one worth externalizing).\n\nSecondly, why externalize these translation operations in particular? As I will now discuss, any motivation for externalizing these particular translation operations must come from the dynamics. Forgoing any dynamical considerations, all we can say about KG1-KG7 is that they concern vectors $\\bm{\\Phi}\\in\\mathbb{R}^L$ (or alternatively functions \\mbox{$\\phi:L\\to\\mathbb{R}$} in $F_L$). Recall that pre-dynamics the set of labels for lattice sites $L$ is uncountable but otherwise unstructured; We might index it using any number of indices we like, \\mbox{$L\\cong\\mathbb{Z}\\cong\\mathbb{Z}^2\\cong\\dots\\cong\\mathbb{Z}^{17}\\cong\\dots$}. Using each of these re-indexings we can grant $\\mathbb{R}^L$ different tensor product structures, \\mbox{$\\mathbb{R}^L\\cong\\mathbb{R}^Z\\cong\\mathbb{R}^Z\\otimes\\mathbb{R}^Z\\cong\\dots$}. In each tensor factor we can define a translation operation $T^\\epsilon$ as in Eq.~\\eqref{TDef}. Thus, the vector space $\\mathbb{R}^L$ supports representations of: the 1D translation group, the 2D translation group, $\\dots$, the 17D translation group, etc. Given all these possibilities, why externalize $T^{\\epsilon_1}_\\text{j}$, $T^{\\epsilon_2}_\\text{n}$, and $T^{\\epsilon_3}_\\text{m}$ in particular? The answer must come from the dynamics.\n\nOne good reason to consider $T^{\\epsilon_1}_\\text{j}$, $T^{\\epsilon_2}_\\text{n}$, and $T^{\\epsilon_3}_\\text{m}$ worthy of externalization is that they are dynamical symmetries of KG4-KG7. However, as mentioned above, something might be a translation operation (and moreover, one worthy of externalizing) even if it's not a dynamical symmetry. Unfortunately, I don't have a perfect rule for how to identify such cases. The best I can offer is to check whether its associated derivative (or some function thereof) appears in the dynamical equations.\n\nRegardless, its sufficiently clear for KG4-KG7 that $T^{\\epsilon_1}_\\text{j}$, $T^{\\epsilon_2}_\\text{n}$, and \n$T^{\\epsilon_3}_\\text{m}$ are among the translation operations worthy of externalizing. Our translation-matching constraint suggests that these ought to correspond (perhaps in a complicated way) to parallel transport on the new spacetime manifold $\\mathcal{M}$. I claim that (at least in this case) this constraint fixes the spacetime manifold $\\mathcal{M}$ up to diffeomorphism.\n\nTo show this I will first pick out at each point \\mbox{$p\\in\\mathcal{M}$} on the manifold three independent directions in the tangent space at $p$. To realize the translation-matching constraint, I demand that differential translation by $T^{\\epsilon_1}_\\text{j}$, $T^{\\epsilon_2}_\\text{n}$, and $T^{\\epsilon_3}_\\text{m}$ is then to be carried out on the manifold as parallel transport in these three directions. Note that this already implies that the dimension of $\\mathcal{M}$ is at least three.\n\nIndeed, in this case we have good reason to take $\\mathcal{M}$ to be exactly three dimensional. To see this, note that \n\\begin{align}\\label{RLCover}\n\\mathbb{R}^L&\\cong \\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\\\\n&=\\text{span}(T^{\\epsilon_1}_\\text{j} T^{\\epsilon_2}_\\text{n} T^{\\epsilon_3}_\\text{m} \\, \\bm{e}_0\\!\\otimes\\bm{e}_0\\!\\otimes\\bm{e}_0\\vert (\\epsilon_1,\\epsilon_2,\\epsilon_3)\\in\\mathbb{R}^3)\n\\end{align}\nThat is, beginning from $\\bm{e}_0\\otimes\\bm{e}_0\\otimes\\bm{e}_0$ these translations cover all of $\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$. That is, they cover all of our state space in the second interpretation. The three translations $T^{\\epsilon_1}_\\text{j}$, $T^{\\epsilon_2}_\\text{n}$, and $T^{\\epsilon_3}_\\text{m}$ are thus already enough to cover every kinematic possibility. We need no further dimensions and so by Occam's razor we may then take the spacetime manifold $\\mathcal{M}$ to be three dimensional. \n\nBut how does the translation-matching constraint (which I have yet to state technically) fix $\\mathcal{M}$ up to diffeomorphism? As I will now discuss, (at least in this case) $\\mathcal{M}$ it is fixed up to diffeomorphism by the group theoretic properties of $T^{\\epsilon_1}_\\text{j}$, $T^{\\epsilon_2}_\\text{n}$, and $T^{\\epsilon_3}_\\text{m}$. Let $G_\\text{trans.}$ be the group formed by all compositions of $T^{\\epsilon_1}_\\text{j}$, $T^{\\epsilon_2}_\\text{n}$, and $T^{\\epsilon_3}_\\text{m}$. Note that the group $G_\\text{trans.}$ is a Lie group, and hence is also a differentiable manifold.\n\nRecall that so far the translation-matching constraint associated to each of our differential translations, a direction in the tangent space of each point $p\\in\\mathcal{M}$. Since $G_\\text{trans.}$ contains nothing but these translations, this gives us a one-to-one correspondence between the algebra $g_\\text{trans.}$ of $G_\\text{trans.}$ and (at least a subspace of) the of the tangent space of each point on the manifold. Let's take these tangent vectors to very smoothly across the manifold. Let $\\mathcal{M}_\\text{trans.}(p)$ be the submanifold of points on the spacetime manifold reachable from $p\\in\\mathcal{M}$ by following these tangent vectors. The algebra $g_\\text{trans.}$ is related to the group $G_\\text{trans.}$ in the same way that this tangent space it related to $\\mathcal{M}_\\text{trans.}(p)$, namely by repeated application of the exponential map. Hence, I take translation-matching to demand that $\\mathcal{M}_\\text{trans.}(p)\\cong G_\\text{trans.}$ as differentiable manifolds for every $p\\in\\mathcal{M}$. \n\nAs discussed above, we have reason to take $\\mathcal{M}$ to be three dimensional. Hence (assuming that $\\mathcal{M}$ is connected) any three dimensional submanifold (e.g. $\\mathcal{M}_\\text{trans.}(p)$) must in fact be the entire manifold. Therefore we have $\\mathcal{M}\\cong G_\\text{trans.}$ as differentiable manifolds.\n\nBut what can we say about this translation group $G_\\text{trans.}$ viewed as a manifold? Notice that these three continuous translation operations comprising $G_\\text{trans.}$ all commute with each other. As such for any combination of them $T_\\text{generic}\\in G_\\text{trans.}$ we always have a unique factorization of the form,\n\\begin{align}\\label{Tgeneric}\nT_\\text{generic} = T^{\\epsilon_1}_\\text{j} \\ T^{\\epsilon_2}_\\text{n} \\ T^{\\epsilon_3}_\\text{m}\n\\end{align}\nMoreover, this factorization represents all of $G_\\text{trans.}$ without redundancy. Indeed, we can smoothly parameterize all of $G_\\text{trans.}$ using parameters \\mbox{$(\\epsilon_1,\\epsilon_2,\\epsilon_3)\\in\\mathbb{R}^3$} in a one-to-one way. That is, $G_\\text{trans.}\\cong\\mathbb{R}^3$ as differentiable manifolds. Therefore, we have $\\mathcal{M}\\cong\\mathbb{R}^3$ as well. Note this means that we have access to a global coordinate system for $\\mathcal{M}$.\n\nThus, from the translation-matching constraint alone we are forced to take $\\mathcal{M}\\cong\\mathbb{R}^3$ for KG4-KG7. For KG1-KG3 the same translation-matching constraint forces us to take $\\mathcal{M}\\cong\\mathbb{R}^2$. Note that in both cases these manifolds satisfy the symmetry-fitting constraint: $\\text{Diff}(\\mathcal{M})$ is large enough to contain $G^\\text{dyn}_\\text{to-be-ext}$.\n\n\\subsection{A Principled Choice of Embedding}\nNow that we have a spacetime manifold selected, we need to somehow embed $\\bm{\\Phi}$ (or equivalently $\\phi_\\ell$) into it. Here I will begin with $\\bm{\\Phi}$ with the approach from $\\phi_\\ell$ being addressed in Sec.~\\ref{SecExtPart2}. The goal in either case is to construct in a principled way from either $\\bm{\\Phi}$ or $\\phi_\\ell$ a new field \\mbox{$\\phi:\\mathcal{M}\\to\\mathbb{R}$} defined over this manifold. \n\nBefore this, however, recall that our move from the first to the second interpretation was mediated by means of a vector-space isomorphism $F_L\\cong\\mathbb{R}^L$ namely Eq.~\\eqref{PhiDef}. Here too, our reinterpretation will be mediated by a vector-space isomorphism.\n\nLet $F(\\mathcal{M})$ be the set of all real scalar functions \\mbox{$f:\\mathcal{M}\\to\\mathbb{R}$}. Notice that this is a vector space as it is closed under addition and scalar multiplication. Our goal here is to in a principle way construct an isomorphism between \\mbox{$\\mathbb{R}^L$} and some vector subspace of \\mbox{$F(\\mathcal{M})$}. Indeed, $F(\\mathcal{M})$ is much larger than $\\mathbb{R}^L$ (having an uncountably infinite dimension). As such we will only ever map $\\mathbb{R}^L$ onto some subset of it, $F\\subset F(\\mathcal{M})$. In order for us to have a vector space isomorphism between $\\mathbb{R}^L$ and $F$ the following conditions must be satisfied. $F$ must be: 1) closed under addition and scalar multiplication and hence a vector space, and 2) countably infinite dimensional and hence isomorphic to $\\mathbb{R}^L$, i.e., \\mbox{$F\\cong\\mathbb{R}^\\mathbb{Z}\\cong\\mathbb{R}^L$}. With the target subset characterized, the embedding of $\\bm{\\Phi}$ onto $\\mathcal{M}$ is then accomplished by picking a vector-space isomorphism $E:\\mathbb{R}^L\\to F$. The embedded field is then gives to us as $\\phi\\coloneqq E(\\bm{\\Phi})$.\n\nIt appears we have a great deal of freedom here. Suppose that, in line with the translation-matching constraint discussed above, we take $\\mathcal{M}\\cong\\mathbb{R}^2$ for KG1-KG3 and $\\mathcal{M}\\cong\\mathbb{R}^3$ for KG4-KG7. In this case, we still have great freedom in picking both the target vector space $F\\subset F(\\mathcal{M})$ and the isomorphism $E:\\mathbb{R}^L\\to F$. However, as I will now discuss, translation-matching also drastically limits these choices as well. Allow me to demonstrate with KG4-KG7.\n\nFirst, allow me to explicitly define a coordinate system for $\\mathcal{M}\\cong G_\\text{trans.}\\cong\\mathbb{R}^3$. As discussed above we already have an explicit smooth parametrization of $G_\\text{trans.}$ via \\mbox{$(\\epsilon_1,\\epsilon_2,\\epsilon_3)\\in\\mathbb{R}^3$}, see Eq.~\\eqref{Tgeneric}. In addition to this, concretely realizing our translation-matching constraint requires us to fix a smooth map from $G_\\text{trans.}$ to $\\mathcal{M}$. Combined these two maps give us a global coordinate system $(\\epsilon_1,\\epsilon_2,\\epsilon_3)\\in\\mathbb{R}^3$ for $\\mathcal{M}$. We can then rescale these to give new coordinates $(t,x,y)\\in\\mathbb{R}^3$ by adding in a length scale $a>0$ with $t=\\epsilon_1 a$, $x=\\epsilon_2 a$, and $y=\\epsilon_3 a$ for some fixed length scale $a$. \n\nIn total this picks out a diffeomorphism \\mbox{$d_\\text{trans.}:\\mathbb{R}^3\\to\\mathcal{M}$} which assigns global coordinates $(t,x,y)\\in\\mathbb{R}^3$ to $\\mathcal{M}$. For much of the next sections I will work with $\\phi$ in these coordinates. Namely I will work with the pull back, $\\phi\\circ d_\\text{trans.}:\\mathbb{R}^3\\to\\mathbb{R}$. One may worry that this choice of coordinates is arbitrary. Indeed, if one changes the smooth map from $G_\\text{trans.}$ to $\\mathcal{M}$ referenced above we end up with a different coordinate system $d_\\text{trans.}$ related to the original by some diffeomorphism. While this is true, no matter how one realizes the translation-matching constraint there will be some coordinate system would result from the above construction. In what follows, nothing depends on how the translation-matching constraint is realized.\n\nA word of warning. The coordinates used above might end up lying on the manifold in a curvilinear way. Crucially, the naive distance and metric structure associated with these coordinates do not automatically have any spaciotemporal significance. At present, our manifold $\\mathcal{M}$ still has no metric. When a metric arises later, it will be due to dynamical considerations, not from any choice of coordinates. A coordinate-independent view of these theories will be given in Sec.~\\ref{SecFullGenCov}.\n\nWith this said, by construction we have the following in correspondences in these coordinate:\n\\begin{flushleft}\\begin{enumerate}\n \\item Action by $T_\\text{j}^\\epsilon$ on $\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$ acts on the manifold $\\mathcal{M}$ in these coordinates as $(t,x,y)\\mapsto(t-\\epsilon\\,a,x,y)$,\n \\item Action by $T_\\text{n}^\\epsilon$ on $\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$ acts on the manifold $\\mathcal{M}$ in these coordinates as $(t,x,y)\\mapsto(t,x-\\epsilon\\,a,y)$,\n \\item Action by $T_\\text{m}^\\epsilon$ on $\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$ acts on the manifold $\\mathcal{M}$ in these coordinates as $(t,x,y)\\mapsto(t,x,y-\\epsilon\\,a)$.\n\\end{enumerate}\\end{flushleft}\nIn particular, matching up differential translations on $\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$ with those on $\\mathcal{M}$ requires\n\\begin{align}\nE(D_\\text{j}\\,\\bm{\\Phi})&=a\\,\\partial_t E(\\bm{\\Phi})\\\\\n\\nonumber\nE(D_\\text{n}\\bm{\\Phi})&=a\\,\\partial_x E(\\bm{\\Phi})\\\\\n\\nonumber\nE(D_\\text{m}\\bm{\\Phi})&=a\\,\\partial_y E(\\bm{\\Phi})\n\\end{align}\nin these coordinates for all $\\bm{\\Phi}\\in\\mathbb{R}^L$.\n\nEvaluating these conditions in the planewave basis for $\\mathbb{R}^L\\cong\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}\\otimes\\mathbb{R}^\\mathbb{Z}$ (namely $\\bm{\\Phi}(\\omega,k_1,k_2)$ for \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$}) we see that $E(\\bm{\\Phi}(\\omega,k_1,k_2))$ must be simultaneously an eigenvector of $\\partial_t$, $\\partial_x$, and $\\partial_y$ with eigenvalues $-\\mathrm{i} \\omega\/a$, $-\\mathrm{i} k_1\/a$, and $-\\mathrm{i} k_2\/a$ respectively. This uniquely picks out the continuum planewaves:\n\\begin{align}\\label{PlaneWaveCont}\n\\text{KG1-KG3:}\\quad&\\phi(t,x;\\omega,k)\n\\coloneqq e^{-\\mathrm{i} \\omega t -\\mathrm{i} k x}\\\\\n\\nonumber\n\\text{KG4-KG7:}\\quad&\\phi(t,x,y;\\omega,k_1,k_2)\n\\coloneqq e^{-\\mathrm{i} \\omega t -\\mathrm{i} k_1 x-\\mathrm{i} k_2 y}.\n\\end{align}\nIn particular, we are forced to take\n\\begin{align}\\label{Eftilde}\nE:\\,&\\bm{\\Phi}(\\omega,k_1,k_2)\\\\\n\\nonumber\n&\\mapsto\\tilde{f}(\\omega,k_1,k_2)\\, \\phi(t,x,y;\\omega\/a,k_1\/a,k_2\/a)\n\\end{align}\nfor some complex function $\\tilde{f}(\\omega,k_1,k_2)\\in\\mathbb{C}$. That is, each discrete planewave must map onto something proportional to the corresponding continuous planewaves (rescaled by $a$)\n\nNote that up to a local rescaling of Fourier space (which recall is a symmetry of the dynamics) our embedding is uniquely fixed by translation-matching. Even without this consideration however, our translation-matching constraint still fixes the vector space $F\\subset F(\\mathcal{M})$ we can embed \\mbox{$\\bm{\\Phi}\\in\\mathbb{R}^L$} onto, note the following. We must have \\mbox{$\\tilde{f}(\\omega,k_1,k_2)\\neq0$} for all \\mbox{$\\omega,k_1,k_2\\in[-\\pi,\\pi]$} otherwise $E$ will not be invertible an hence not an isomorphism. Therefore $F$ is fixed as,\n\\begin{align}\n&F=\\text{span}(E(\\bm{\\Phi}(\\omega,k_1,k_2))\\vert \\omega,k_1,k_2\\in[-\\pi,\\pi])\\\\\n\\nonumber\n&=\\!\\text{span}(\\tilde{f}(\\omega,k_1,k_2)\\,\\phi(t,x,y;\\frac{\\omega}{a},\\frac{k_1}{a},\\frac{k_2}{a})\\vert \\omega,k_1,k_2\\in\\![-\\pi,\\pi])\\\\\n\\nonumber\n&=\\text{span}(\\phi(t,x,y;\\frac{\\omega}{a},\\frac{k_1}{a},\\frac{k_2}{a})\\vert \\omega,k_1,k_2\\in[-\\pi,\\pi])\\\\\n\\nonumber\n&=\\text{span}(\\phi(t,x,y;\\omega,k_1,k_2)\\vert \\omega,k_1,k_2\\in[-\\pi\/a,\\pi\/a]).\n\\end{align}\nIn light of this, let us define\n\\begin{align}\n&\\text{KG1-KG3:}\\\\\n\\nonumber\n&F^K\\coloneqq\\text{span}(\\phi(t,x;\\omega,k)\\vert \\omega,k\\in[-K,K])\\\\\n&\\text{KG4-KG7:}\\\\\n\\nonumber\n&F^K\\coloneqq\\text{span}(\\phi(t,x,y;\\omega,k_1,k_2)\\vert \\omega,k_1,k_2\\in[-K,K])\n\\end{align}\nwhere $K=\\pi\/a$. These are the spaces of bandlimited functions with bandwidth $\\omega,k\\in[-K,K]$ and \\mbox{$\\omega,k_1,k_2\\in[-K,K]$} respectively. Thus, demanding translation-matching we are forced to map $\\bm{\\Phi}\\in\\mathbb{R}^L$ onto some bandlimited $\\phi(t,x,y)\\in F=F^K$ (at least in the coordinates $d_\\text{trans.}$).\n\nThe next section will discuss in detail some remarkable properties of bandlimited functions, namely their sampling property. Before that however, let's find an interpretation for the $\\tilde{f}(\\omega,k_1,k_2)$ function appearing in Eq.~\\eqref{Eftilde}. First note that applying $E$ to the basis vector $\\bm{e}_0\\otimes\\bm{e}_0\\otimes\\bm{e}_0$ we have\n\\begin{align}\\label{Embedf}\nE(\\bm{e}_0\\otimes\\bm{e}_0\\otimes\\bm{e}_0)=f\\!\\left(\\frac{t}{a},\\frac{x}{a},\\frac{y}{a}\\right)\n\\end{align}\nwhere $f(t,x,y)$ is the inverse Fourier transform of $\\tilde{f}(\\omega,k_1,k_2)$. Next note that by applying $T_\\text{j}^\\epsilon$, $T_\\text{n}^\\epsilon$, and $T_\\text{m}^\\epsilon$ with integer arguments we can get from $\\bm{e}_0\\otimes\\bm{e}_0\\otimes\\bm{e}_0$ to any other basis vector, $\\bm{e}_j\\otimes\\bm{e}_n\\otimes\\bm{e}_m$. In these coordinates this means,\n\\begin{align}\n&E(\\bm{e}_j\\otimes\\bm{e}_n\\otimes\\bm{e}_m)=f\\left(\\frac{t}{a}-j,\\frac{x}{a}-n,\\frac{y}{a}-m\\right)\n\\end{align}\nThus, we can understand a choice of $\\tilde{f}$ as picking a profile $f(t,x,y)\\in F^\\pi$. A translated and rescaled copy of this profile is then associated with each basis vector.\n\nTo review: Our translation-matching considerations have greatly constrained our choice of both the spacetime manifold $\\mathcal{M}$ and the embedding of $\\bm{\\Phi}$ onto $F(\\mathcal{M})$. In particular, $\\mathcal{M}$ was forced to be diffeomorphic to $\\mathbb{R}^2$ for KG1-KG3 and to $\\mathbb{R}^3$ for KG4-KG7. Moreover, the vector space $F\\subset F(\\mathcal{M})$ which we map $\\bm{\\Phi}\\in\\mathbb{R}^L$ into is forced to be the space of bandlimited functions with some bandwidth (i.e., bandlimited in coordinates $d_\\text{trans.}$ with $\\omega,k_1,k_2\\in[-K,K]$).\n\nBeyond this, our only freedom left is in picking a profile $f$ to associate with each basis vector $\\bm{e}_j\\otimes\\bm{e}_n\\otimes\\bm{e}_m$. In Sec.~\\ref{SecExtPart2} I will motivate a principled choice for $f$. Before that however, let's discuss bandlimited function in some detail.\n\n\\section{Brief Review of Bandlimited Functions and Nyquist-Shannon Sampling Theory}\\label{SecSamplingTheory}\n The previous section has given us reason to care about bandlimited functions. A bandlimited function is one whose Fourier transform has compact support. The bandwidth of such a function is the extent of its support in Fourier space. As I will now discuss, such functions have a remarkable sampling property: they can be exactly reconstructed knowing only their values at a (sufficiently dense) set of sample points. The study of such functions constitutes Nyquist-Shannon sampling theory. For a selection of introductory texts on sampling theory see ~\\cite{GARCIA200263,SamplingTutorial,UnserM2000SyaS}.\n\nTo introduce the topic I will at first restrict our attention to the one-dimensional case with uniform sample lattice before generalizing to higher dimensions and non-uniform samplings later on.\n\n\\subsection{One Dimension Uniform Sample Lattices}\\label{Sec1DUniform}\nConsider a generic bandlimited function, $f_\\text{B}(x)$, with a bandwidth of $K$. That is, a function $f_\\text{B}(x)$ such that its Fourier transform,\n\\begin{align}\n\\mathcal{F}[f_\\text{B}(x)](k)\\coloneqq\\int_{-\\infty}^\\infty f_\\text{B}(x) \\, e^{-\\mathrm{i} k x} \\d x,\n\\end{align}\nhas support only for wavenumbers $\\vert k\\vert< K$.\n\nSuppose that we know the value of $f_\\text{B}(x)$ only at the regularly spaced sample points, $x_n=n\\,a+b$, with some spacing, \\mbox{$0\\leq a\\leq a^*\\coloneqq\\pi\/K$}, and offset, $b\\in\\mathbb{R}$. Let \\mbox{$f_n=f_\\text{B}(x_n)$} be these sample values. Having only the discrete sample data, $\\{(x_n,f_n)\\}_{n\\in\\mathbb{Z}}$, how well can we approximate the function? \n\nThe Nyquist-Shannon sampling theorem~\\cite{ShannonOriginal} tells us that from this data we can reconstruct $f_\\text{B}$ exactly everywhere! That is, from this discrete data, $\\{(x_n,f_n)\\}_{n\\in\\mathbb{Z}}$, we can determine everything about the function $f_\\text{B}$ everywhere. In particular, the following reconstruction is exact, \n\\begin{align}\\label{SincRecon}\nf_\\text{B}(x) \n= S_n\\!\\left(\\frac{x-b}{a}\\right) f_n,\n\\end{align}\nwhere\n\\begin{align}\nS(y)=\\frac{\\sin(\\pi y)}{\\pi y}, \\quad\\text{and}\\quad\nS_n(y)=S(y-n), \n\\end{align}\nare the normalized and shifted sinc functions. Note that $S_n(m)=\\delta_{nm}$ for integers $n$ and $m$. Moreover, note that each $S_n(x)$ is both $L_1$ and $L_2$ normalized and that taken together the set $\\{S_n(x)\\}_{n\\in\\mathbb{Z}}$ forms an orthonormal basis with respects to the $L_2$ inner product. The fact that any bandlimited function can be reconstructed in this way is equivalent to the fact that this orthonormal basis spans the space of bandlimited functions with bandwidth of $K=\\pi$.\n\n\\begin{figure}\n\\includegraphics[width=0.4\\textwidth]{PaperFigures\/Fig1DSamplesa.pdf}\n\\includegraphics[width=0.4\\textwidth]{PaperFigures\/Fig1DSamplesb.pdf}\n\\includegraphics[width=0.4\\textwidth]{PaperFigures\/Fig1DSamplesc.pdf}\n\\includegraphics[width=0.4\\textwidth]{PaperFigures\/Fig1DSamplesd.pdf}\n\\includegraphics[width=0.4\\textwidth]{PaperFigures\/Fig1DSamplese.pdf}\n\\caption{Several different (but completely equivalent) graphical representations of the bandlimited function \\mbox{$f_\\text{B}(x)=1+S(x-1\/2)+x\\,S(x\/2)^2$} with bandwidth of $K=\\pi$ and consequently a critical spacing of $a^*=\\pi\/K=1$. Subfigure a) shows the function values for all $x$. b) shows the values of $f_\\text{B}$ at $x_n=n\/2$. Since $1\/20$. As such, we can represent $f_\\text{B}(x,y)$ with a (sufficiently dense) uniform sample lattice in both the $x$ and $y$ directions. That is, we can represent any bandlimited $f_\\text{B}(x,y)$ in terms of its sample values on a sufficiently dense square lattice. Once we have such a uniform sampling, the reasoning carried out above applies unchanged. We can represent $f_\\text{B}(x,y)$ on any sufficiently dense non-uniform lattice.\n\nFor a concrete example consider the bandlimited function shown shown in Fig.~\\ref{Fig2DSamples}a), namely,\n\\begin{align}\\label{J1}\nf_\\text{B}(x,y)=J_1(\\pi\\,r)\/(\\pi\\,r) \n\\end{align}\nwhere $J_1$ is the first Bessel function and $r=\\sqrt{x^2+y^2}$. This function is bandlimited with $\\sqrt{k_x^2+k_y^2}