THE KPZ FIXED POINT KONSTANTIN MATETSKI, JEREMY QUASTEL, AND DANIEL REMENIK ABSTRACT. An explicit Fredholm determinant formula is derived for the multipoint distribution of the height function of the totally asymmetric simple exclusion process with arbitrary initial condition. The method is by solving the biorthogonal ensemble/non-intersecting path representation found by [Sas05; BFPS07]. The resulting kernel involves transition probabilities of a random walk forced to hit a curve defined by the initial data. In the KPZ 1:2:3 scaling limit the formula leads in a transparent way to a Fredholm determinant formula, in terms of analogous kernels based on Brownian motion, for the transition probabilities of the scaling invariant Markov process at the centre of the KPZ universality class. The formula readily reproduces known special self-similar solutions such as the Airy1 and Airy2 processes. The invariant Markov process takes values in real valued functions which look locally like Brownian motion, and is Hölder 1/3- in time. arXiv:1701.00018v1 [math.PR] 30 Dec 2016 CONTENTS 1. The KPZ universality class 1 2. TASEP 3 2.1. Biorthogonal ensembles 4 2.2. TASEP kernel as a transition probability with hitting 7 2.3. Formulas for TASEP with general initial data 8 3. 1:2:3 scaling limit 10 4. The invariant Markov process 15 4.1. Fixed point formula 15 4.2. Markov property 16 4.3. Regularity and local Brownian behavior 17 4.4. Airy processes 17 4.5. Symmetries and variational formulas 19 4.6. Regularity in time 20 4.7. Equilibrium space-time covariance 20 Appendix A. Path integral formulas 21 Appendix B. Trace class estimates 23 Appendix C. Regularity 29 References 31 1. THE KPZ UNIVERSALITY CLASS All models in the one dimensional Kardar-Parisi-Zhang (KPZ) universality class (random growth models, last passage and directed polymers, random stirred fluids) have an analogue of the height function h(t, x) (free energy, integrated velocity) which is conjectured to converge at large time and Date: January 3, 2017. This is a preliminary version and will be updated; we welcome comments from readers. 1 THE KPZ FIXED POINT 2 length scales ( 0), under the KPZ 1:2:3 scaling 1/2h(-3/2t, -1x) - Ct, (1.1) to a universal fluctuating field h(t, x) which does not depend on the particular model, but does depend on the initial data class. Since many of the models are Markovian, the invariant limit process, the KPZ fixed point, will be as well. The purpose of this article is to describe this Markov process, and how it arises from certain microscopic models. The KPZ fixed point should not be confused with the Kardar-Parisi-Zhang equation [KPZ86], th = (xh)2 + x2h + 1/2 (1.2) with a space-time white noise, which is a canonical continuum equation for random growth, lending its name to the class. One can think of the space of models in the class as having a trivial, Gaussian fixed point, the Edwards-Wilkinson fixed point, given by (1.2) with = 0 and the 1:2:4 scaling 1/2h(-2t, -1x) - Ct, and the non-trivial KPZ fixed point, given by (1.2) with = 0. The KPZ equation is just one of many models, but it plays a distinguished role as the (conjecturally) unique heteroclinic orbit between the two fixed points. The KPZ equation can be obtained from microscopic models in the weakly asymmetric or intermediate disorder limits [BG97; AKQ14; MFQR17; CT15; CN16; CTS16] (which are not equivalent, see [HQ15]). This is the weak KPZ universality conjecture. However, the KPZ equation is not invariant under the KPZ 1:2:3 scaling (1.1), which is expected to send it, along with all other models in the class, to the true universal fixed point. In modelling, for example, edges of bacterial colonies, forest fires, spread of genes, the non-linearities or noise are often not weak, and it is really the fixed point that should be used in approximations and not the KPZ equation. However, progress has been hampered by a complete lack of understanding of the time evolution of the fixed point itself. Essentially all one had was a few special self-similar solutions, the Airy processes. Under the KPZ 1:2:3 scaling (1.1) the coefficients of (1.2) transform as 1/2. A naive guess would then be that the fixed point is nothing but the vanishing viscosity ( 0) solution of the Hamilton-Jacobi equation th = (xh)2 + x2h given by Hopf's formula h(t, x) = sup{- y t (x - y)2 + h0(y)}. It is not: One of the key features of the class is a stationary solution consisting of (non-trivially) time dependent Brownian motion height functions (or discrete versions). But Brownian motions are not invariant for Hopf's formula (see [FM00] for the computation). Our story has another parallel in the dispersionless limit of KdV ( 0 in) th = (xh)2 + x3h (in integrated form). Brownian motions are invariant for all , at least in the periodic case [QV08]. But as far as we are aware, the zero dispersion limit has only been done on a case by case basis, with no general formulas. All of these lead, presumably, to various weak solutions of the pure non-linear evolution th = (xh)2, which is, of course, ill-posed. Our fixed point is also given by a variational formula (see Theorem 4.10) involving a residual forcing noise, the Airy sheet. But, unfortunately, our techniques do not allow us to characterize this noise. Instead, we obtain a complete description of the Markov field h(t, x) itself through the exact calculation of its transition probabilities (see Theorem 4.1). The strong KPZ universality conjecture (still wide open) is that this fixed point is the limit under the scaling (1.1) for any model in the class, loosely characterized by having: 1. Local dynamics; 2. Smoothing mechanism; 3. Slope dependent growth rate (lateral growth); 4. Space-time random forcing with rapid decay of correlations. Universal fixed points have become a theme in probability and statistical physics in recent years. 4d, SLE, Liouville quantum gravity, the Brownian map, the Brownian web, and the continuum random tree have offered asymptotic descriptions for huge classes of models. In general, these have been obtained as non-linear transformations of Brownian motions or Gaussian free fields, and their description relies to a THE KPZ FIXED POINT 3 large degree on symmetry. In the case of 4d, the main tool is perturbation theory. Even the recent theory of regularity structures [Hai14], which makes sense of the KPZ equation (1.2), does so by treating the non-linear term as a kind of perturbation of the linear equation. In our case, we have a non-perturbative two-dimensional field theory with a skew symmetry, and a solution should not in principle even be expected. What saves us is the one-dimensionality of the fixed time problem, and the fact that several discrete models in the class have an explicit description using non-intersecting paths. Here we work with TASEP, obtaining a complete description of the transition probabilities in a form which allows us to pass transparently to the 1:2:3 scaling limit1. In a sense, a recipe for the solution of TASEP has existed since the work of [Sas05], who discovered a highly non-obvious representation in terms of non-intersecting paths which in turn can be studied using the structure of biorthogonal ensembles [BFPS07]. However, the biorthogonalization was only implicit, and one had to rely on exact solutions for a couple of special initial conditions to obtain the asymptotic Tracy-Widom distributions FGUE and FGOE [TW94; TW96] and the Baik-Rains distribution FBR [BR01], and their spatial versions, the Airy processes [Joh00; Joh03; Sas05; BFPS07; BFP07; BFP10]. In this article, motivated by the probabilistic interpretation of the path integral forms of the kernels in the Fredholm determinants, and exploiting the skew time reversibility, we are able to obtain a general formula in which the TASEP kernel is given by a transition probability of a random walk forced to hit the initial data. We end this introduction with an outline of the paper and a brief summary of our results. Section 2.1 recalls and solves the biorthogonal representation of TASEP, motivated by the path integral representation, which is derived in the form we need it in Appendix A.2. The biorthogonal functions appearing in the resulting Fredholm determinants are then recognized as hitting probabilities in Section 2.2, which allows us to express the kernels in terms of expectations of functionals involving a random walk forced to hit the initial data. The determinantal formulas for TASEP with arbitrary initial conditions are in Theorem 2.6. In Section 3, we pass to the KPZ 1:2:3 scaling limit to obtain determinantal formulas for transition probabilities of the KPZ fixed point. For this purpose it turned out to be easier to use formulas for right-finite initial TASEP data. But since we have exact formulas, we can obtain a very strong estimate (Lemma 3.2) on the propagation speed of information which allows us to show there is no loss of generality in doing so. Section 4 opens with the general formula for the transition probabilities of the KPZ fixed point, Theorem 4.1; readers mostly interested in the physical implications may wish to skip directly there. We then work in Section 4.2 to show that the Chapman-Kolmogorov equations hold. This is done by obtaining a uniform bound on the local Hölder < 1/2 norm of the approximating Markov fields. The proof is in Appendix C. The rest of Section 4 gives the key properties of the fixed point: regularity in space and time and local Brownian behavior, various symmetries, variational formulas in terms of the Airy sheet, and equilibrium space-time covariance; we also show how to recover some of the classical Airy processes from our formulas. Sections 3 and 4 are done at the level of pointwise convergence of kernels, skipping moreover some of the details. The convergence of the kernels is upgraded to trace class in Appendix B, where the remaining details are filled in. So, in a sense, everything follows once one is able to explictly biorthogonalize TASEP. We begin there. 2. TASEP The totally asymmetric simple exclusion process (TASEP) consists of particles with positions · · · < Xt(2) < Xt(1) < Xt(0) < Xt(-1) < Xt(-2) < · · · on Z {-, } performing totally asymmetric nearest neighbour random walks with exclusion: Each particle independently attempts jumps to the neighbouring site to the right at rate 1, the jump being allowed only if that site is unoccupied (see [Lig85] for the non-trivial fact that the process with an infinite number of particles makes sense). Placing a necessarily infinite number of particles at ± allows for left- or right-finite data with no change of notation, the particles at ± playing no role in the dynamics. We follow the standard 1The method works for several variants of TASEP which also have a representation through biorthogonal ensembles, which will appear in the updated version of this article. THE KPZ FIXED POINT 4 practice of ordering particles from the right; for right-finite data the rightmost particle is labelled 1. Let Xt-1(u) = min{k Z : Xt(k) u} denote the label of the rightmost particle which sits to the left of, or at, u at time t. The TASEP height function associated to Xt is given for z Z by ht(z) = -2 Xt-1(z - 1) - X0-1(-1) - z, (2.1) which fixes h0(0) = 0. We will also choose the frame of reference X0-1(-1) = 1, i.e. the particle labeled 1 is initially the rightmost in Z<0. The height function is a random walk path ht(z + 1) = ht(z) + ^t(z) with ^t(z) = 1 if there is a particle at z at time t and -1 if there is no particle at z at time t. The dynamics of ht is that local max's become local min's at rate 1; i.e. if ht(z) = ht(z ± 1) + 1 then ht(z) ht(z) - 2 at rate 1, the rest of the height function remaining unchanged. We can also easily extend the height function to a continuous function of x R by linearly interpolating between the integer points. 2.1. Biorthogonal ensembles. TASEP was first solved by Schütz [Sch97] using Bethe ansatz. He showed that the transition probability for N particles has a determinantal form P(Xt(1) = x1, . . . , Xt(N ) = xN ) = det(Fi-j(xN+1-i - X0(N + 1 - j), t))1i,jN (2.2) with (-1)n Fn(x, t) = 2i 0,1 dw w (1 - w)-n wx-n et(w-1), where 0,1 is any simple loop oriented anticlockwise which includes w = 0 and w = 1. To mesh with our convention of infinitely many particles, we can place particles X0(j), j 0 at and X0(j), j > N at -. Remarkable as it is, this formula is not conducive to asymptotic analysis where we want to consider the later positions of M N of the particles. This was overcome by [Sas05; BFPS07] who were able to reinterpret the integration of (2.2) over the excess variables as a kind of non-intersecting line ensemble, and hence the desired probabilities could be obtained from a biorthogonalization problem, which we describe next. First for a fixed vector a RM and indices n1 < . . . < nM we introduce the functions a(nj, x) = 1x>aj , ¯a(nj, x) = 1xaj , which also regard as multiplication operators acting on the space 2({n1, . . . , nM } × Z) (and later on L2({t1, . . . , tM } × R)). We will use the same notation if a is a scalar, writing a(x) = 1 - ¯a(x) = 1x>a. Theorem 2.1 ([BFPS07]). Suppose that TASEP starts with particles labeled 1, 2, . . . (so that, in particular, there is a rightmost particle)2,3 and let 1 n1 < n2 < · · · < nM N . Then for t > 0 we have where P(Xt(nj) aj, j = 1, . . . , M ) = det(I - ¯aKt¯a) 2({n1,...,nM }×Z) nj Kt(ni, xi; nj, xj) = -Qnj-ni (xi, xj) + nnii-k(xi)nnjj-k(xj ), k=1 (2.3) (2.4) 2We are assuming here that X0(j) < for all j 1; particles at - are allowed. 3The [BFPS07] result is stated only for initial conditions with finitely many particles, but the extension to right-finite (infinite) initial conditions is straightforward because, given fixed indices n1 < n2 < · · · < nM , the distribution of Xt(n1), . . . , Xt(nM ) does not depend on the initial positions of the particles with indices beyond nM . THE KPZ FIXED POINT 5 and where4 1 Q(x, y) = 2x-y 1x>y and nk (x) = 1 2i dw (1 - w)k et(w-1), 0 2x-X0(n-k)wx+k+1-X0(n-k) (2.5) where 0 is any simple loop, anticlockwise oriented, which includes the pole at w = 0 but not the one at w = 1. The functions nk (x), k = 0, . . . , n - 1, are defined implicitly by (1) The biorthogonality relation xZ nk (x)n(x) = 1k= ; (2) 2-xnk (x) is a polynomial of degree at most n - 1 in x for each k. The initial data appear in a simple way in the nk , which can be computed explicitly. Qm is easy, Qm(x, y) = 1 2x-y x-y-1 m-1 1xy+m, and moreover Q and Qm are invertible: Q-1(x, y) = 2 · 1x=y-1 - 1x=y, Q-m(x, y) = (-1)y-x+m2y-x m . y-x (2.6) It is not hard to check [BFPS07, Eq. 3.22] that for all m, n Z, Qn-mnn-k = mm-k. In particular, nk = Q-kn0-k, while by Cauchy's residue theorem we have n0 = RX0(n), where y(x) = 1x=y and 1 R(x, y) = 2i dw 0 e-t(1-w) 2x-y wx-y+1 = e-t tx-y 2x-y(x - y)! 1xy . R is also invertible, with R-1(x, y) = 1 2i dw 0 et(1-w) 2x-y wx-y+1 = et (-t)x-y 2x-y(x - y)! 1xy . Q and R commute, because Q(x, y) and R(x, y) only depend on x - y. So (2.7) nk = RQ-kX0(n-k). (2.8) The nk , on the other hand, are defined only implicitly through 1 and 2. Only for a few special cases of initial data (step, see e.g. [Fer15]; and periodic [BFPS07; BFP07; BFS08]) were they known, and hence only for those cases asymptotics could be performed, leading to the Tracy-Widom FGUE and FGOE one-point distributions, and then later to the Airy processes for multipoint distributions. We are now going to solve for the nk for any initial data. Let us explain how this can be done starting just from the solution for step initial data X0(i) = -i, i 1. In addition to the extended kernel formula (2.3), one has a path integral formula (see Appendix A.2 for the proof), det I - Kt(nm)(I - Qn1-nm a1 Qn2-n1 a2 · · · Qnm-nm-1 am ) L2(R), (2.9) where Kt(n) = Kt(n, ·; n, ·). (2.10) Such formulas were first obtained in [PS02] for the Airy2 process (see [PS11] for the proof), and later extended to the Airy1 process in [QR13a] and then to a very wide class of processes in [BCR15]. The key is to recognize the kernel Q(x, y) as the transition probabilities of a random walk (which is why we conjugated the [BFPS07] kernel by 2x) and then a1 Qn2-n1 a2 · · · Qnm-nm-1 am (x, y) as the probability that this walk goes from x to y in nm - n1 steps, staying above a1 at time n1, above a2 at time n2, etc. Next we use the skew time reversibility of TASEP, which is most easily stated in terms of the height function, Pf (ht(x) g(x), x Z) = P-g(ht(x) -f (x), x Z) , (2.11) 4We have conjugated the kernel Kt from [BFPS07] by 2x for convenience. The additional X0(n - k) in the power of 2 in the nk 's is also for convenience and is allowed because it just means that the nk 's have to be multiplied by 2X0(n-k). THE KPZ FIXED POINT 6 the subscript indicating the initial data. In other words, the height function evolving backwards in time is indistiguishable from minus the height function. Now suppose we have the solution (2.4) for step initial data centered at x0, which means h0 is the peak -|x - x0|. The multipoint distribution at time t is given by (2.9), but we can use (2.11) to reinterpret it as the one point distribution at time (t, x0), starting from a series of peaks. The multipoint distributions can then be obtained by extending the resulting kernel in the usual way, as in (2.4) (see also (A.1)). One can obtain the general formula and then try to justify proceeding in this fashion. But, in fact, it is easier to use this line of reasoning to simply guess the formula, which can then be checked from Theorem 2.1. This gives us our key result. Theorem 2.2. Fix 0 k < n and consider particles at X0(1) > X0(2) > · · · > X0(n). Let hnk ( , z) be the unique solution to the initial­boundary value problem for the backwards heat equation (Q )-1 hnk ( , z) = hnk ( + 1, z) hnk (k, z) = 2z-X0(n-k) hnk ( , X0(n - k)) = 0 < k, z Z; z Z; < k. Then nk (z) = (R)-1hnk (0, ·)(z) = hnk (0, y)R-1(y, z). yZ Here Q(x, y) = Q(y, x) is the kernel of the adjoint of Q (and likewise for R). (2.12a) (2.12b) (2.12c) Remark 2.3. It is not true that Qhnk ( + 1, z) = hnk ( , z). In fact, in general Qhnk (k, z) is divergent. Proof. The existence and uniqueness is an elementary consequence of the fact that the dimension of ker(Q)-1 is 1, and it consists of the function 2z, which allows us to march forwards from the initial condition hnk (k, z) = 2z-X0(n-k) uniquely solving the boundary value problem hnk ( , X0(n - k)) = 0 at each step. We next check the biorthogonality. n(z)nk (z) = R(z, z1)Q- (z1, X0(n - ))hnk (0, z2)R-1(z2, z) zZ z,z1,z2Z = Q- (z, X0(n - ))hnk (0, z) = (Q)- hnk (0, X0(n - )). zZ For k, we use the boundary condition hnk ( , X0(n - k)) = 1 =k, which is both (2.12b) and (2.12c), to get (Q)- hnk (0, X0(n - )) = hnk ( , X0(n - )) = 1k= . For > k, we use (2.12a) and 2z ker (Q)-1 (Q)- hnk (0, X0(n - )) = (Q)-( -k-1)(Q)-1hnk (k, X0(n - )) = 0. Finally, we show that 2-xnk (x) is a polynomial of degree at most k in x. We have 2-xnk (x) = 2-x et (-t)y-x 2y-x(y - x)! h(0n,k)(y) = et (-t)y y! 2-(x+y)h0(n,k)(x + y). yx y0 We will show that 2-xhnk (0, x) is a polynomial of degree at most k in x. From this it follows that y0 (-t)y y! 2-(x+y)hnk (0, x + y) = e-t pk (x) for some polynomial pk of degree at most k, and thus we get 2-xnk (x) = pk(x). To see that 2-xhnk (0, x) is a polynomial of degree at most k, we proceed by induction. Note first that, by (2.12b), 2-xhnk (k, x) is a polynomial of degree 0. Assume now that 2-xhnk ( , x) is a polynomial of degree at most k - for some 0 < k. By (2.12a) and (2.6) we have 2-xhnk ( , x) = 2-x(Q)-1hnk ( - 1, x) = 2-(x-1)hnk ( - 1, x - 1) - 2-xhnk ( - 1, x), which implies that 2-xhnk ( - 1, x) = hnk ( , X0(n - k)) - x j=X0(n-k)+1 2-j hnk ( , j), which (using (2.12b) and (2.12c)) is a polynomial of degree at most k - + 1 by the inductive hypothesis. THE KPZ FIXED POINT 7 2.2. TASEP kernel as a transition probability with hitting. We will restrict for a while to the single time kernel Kt(n) defined in (2.10). The multi-time kernel can then be recovered as (see (A.5)) Kt(ni, ·; nj , ·) = -Qnj-ni 1ni X0(n - m)}, with the convention that min = . Then for z X0(n - ) we have hnk ( , z) = PB -1 =z ,n = k , which can be proved by checking that (Q)-1hnk ( , ·)¯X0(n- ) = hnk ( + 1, ·)¯X0(n- -1). From the memoryless property of the geometric distribution we have for all z X0(n - k) that PB- 1=z 0,n = k, Bk = y = 2X0(n-k)-yPB- 1=z 0,n = k , and as a consequence we get, for z2 X0(n), n-1 G0,n(z1, z2) = PB- 1=z2 0,n = k (Q)n-k(X0(n - k), z1) k=0 n-1 = PB- 1=z2 0,n = k, Bk = z (Q)n-k-1(z, z1) k=0 z>X0(n-k) = PB- 1=z2 0,n < n, Bn-1 = z1 , (2.16) which is the probability for the walk starting at z2 at time -1 to end up at z1 after n steps, having hit the curve X0(n - m) m=0,...,n-1 in between. The next step is to obtain an expression along the lines of (2.16) which holds for all z2, and not just z2 X0(n). We begin by observing that for each fixed y1, 2-y2Qn(y1, y2) extends in y2 to a polynomial 2-y2Q(n)(y1, y2) of degree n - 1 with Q(n)(y1, y2) = 1 2i (1 + w)y1-y2-1 dw 0 2y1-y2 wn = (y1 - y2 - 1)n-1 2y1-y2 (n - 1)! , (2.17) where (x)k = x(x - 1) · · · (x - k + 1) for k > 0 and (x)0 = 1 is the Pochhammer symbol. Note that Q(n)(y1, y2) = Qn(y1, y2), y1 - y2 1. (2.18) Using (2.6) and (2.17), we have Q-1Q(n) = Q(n)Q-1 = Q(n-1) for n > 1, but Q-1Q(1) = Q(1)Q-1 = 0. Note also that Q(n)Q(m) is divergent, so the Q(n) are no longer a group like Qn. Let = min{m 0 : Bm > X0(m + 1)}, (2.19) where Bm is now a random walk with transition matrix Q (that is, Bm has Geom[ 1 2 ] jumps strictly to the left). Using this stopping time and the extension of Qm we obtain: 5We use the notation Bm to distinguish it from the walk with transition probabilities Q which will appear later. THE KPZ FIXED POINT 8 Lemma 2.4. For all z1, z2 Z we have G0,n(z1, z2) = 1z1>X0(1)Q(n)(z1, z2) + 1z1X0(1)EB0=z1 Q(n- )(B , z2)1 X0(k+1) (2.20) The last expectation is straightforward to compute if z1 > X0(1), and we get G0,n(z1, z2) = 1z1>X0(1)Qn(z1, z2) + 1z1X0(1)EB0=z1 Qn- B , z2 1 X0(1)Q(n)(z1, z2) + 1z1X0(1)EB0=z1 Q(n- ) B , z2 1 0, where P(Xt(nj) aj, j = 1, . . . , M ) = det(I - ¯aKt¯a) 2({n1,...,nM }×Z) , (2.24) Kt(ni, ·; nj , ·) = -Qnj-ni 1ni k - 1 ¯X0N (k-N-1)QX0N (k-N)Q(n+N+1-k) (y, z2) = n-1 k=-N yZ PB0=z1 Bk+N = y, N > k + N ¯X0N (k)QX0N (k+1)Q(n-k) (y, z2). The probability in the above sum equals PB0=y Bk+N = z1, k,N > k + N , where k,N is now the hitting time by Bm of the epigraph of X0N (k - m) m=0,...,k+N , and is in turn given by (Q)k+N (y, z1) - k+N =0 y Z PB0=y k,N = , B = y (Q)k+N- (y , z1). Now we apply Q-N-1 on the z1 variable and then take N to deduce the formula. Example 2.7. (Step initial data) Consider TASEP with step initial data, i.e. X0(i) = -i for i 1. If we start the random walk in (2.23) from B0 = z1 below the curve, i.e. z1 < 0, then the random walk clearly never hits the epigraph. Hence, S¯te,pni(X0) 0 and the last term in (2.25) vanishes. For the second term in (2.25) we have (St,-ni )X0(1)St,nj (z1, z2) = 1 (2i)2 dw 0 (1 - w)ni (1 - v)nj+z2 et(w+v-1) dv 0 2z1-z2 wni+z1+1vnj , 1-v-w 6This formula will not be used in the sequel, so the reader may choose to skip the proof. THE KPZ FIXED POINT 10 which is exactly the formula previously derived in the literature (see e.g. [Fer15, Eq. 82]). Example 2.8. (Periodic initial data) Consider now TASEP with the (finite) periodic initial data X0(i) = 2(N -i) for i = 1, . . . , 2N . For simplicity we will compute only Kt(n). We start by computing S¯te,pni(X0), and proceed formally. Observe that eBm-m() m0, with () = - log(2e - 1) the logarithm of the moment generating function of a negative Geom[ 1 2 ] random variable, is a martingale. Thus if z1 2(N - 1), EB0=z1[eB -()] = ez1. But it is easy to see from the definition of X0 that B is necessarily 2(N - ) +1. Using this and choosing = log(e + e2 - e) leads to EB0=z1 [e- ] = e(z1-2N-1) log(e+ e2-e). Formally inverting the moment generating function gives PB0=z1 ( = k) = 1 2i 0 d ke(z1-2N -1) log(+ 2-). From this we compute, for z1 2(N - 1), that S¯te,pni(X0)(z1, z2) equals 1 (2i)2 dw du 1 2z1-z2 et(w-1/2) (1 - w)z2-2N +n+1 wn-1 (1 - u)2N -z1 (1-w)w (1-u)u n-1 -1 w(1 - w) - u(1 - u) 2u - u 1 , where we have changed variables - (4u(1 - u))-1. From this we may compute the product St,-n)2(N-1)S¯te,pni(X0)(z1, z2), which equals 1 (2i)3 dv dw dw 1 2z1-z2 et(w+v-1) (1 - v)n vn+1+z1 (1 - w)z2-2N +n+1 wn-1 v2N+2 2u - 1 (1-w)w (1-u)u n-1 -1 × . u + v - 1 u w(1 - w) - u(1 - u) Consider separately the two terms coming from the difference in the numerator of the last fraction. Computing the residue at v = 1 - u for the first term leads exactly to the kernel in [BFPS07, Eq. 4.11]. The other term is treated similarly, and it is not hard to check that it cancels with the other summand in (2.25), (St,-n)X0(1)St,n(z1, z2). 3. 1:2:3 SCALING LIMIT For each > 0 the 1:2:3 rescaled height function is h(t, x) = 1/2 h-3/2t(2-1x) + 1 2 -3/2t . (3.1) Remark 3.1. The KPZ fixed point has one free parameter7, corresponding to in (1.2). Our choice of the height function moving downwards corresponds to setting > 0. The scaling of space by the factor 2 in (3.1) corresponds to the choice || = 1/2. Assume that we have initial data X0 chosen to depend on in such a way that8 h0 = lim h(0, ·). 0 (3.2) Because the X0(k) are in reverse order, and because of the inversion (2.1), this is equivalent to 1/2 X0(-1x) + 2-1x - 1 --- -h0(-x). 0 (3.3) 7[JG15] has recently conjectured that the KPZ fixed point is given by th = (xh)2 - (-x2)3/2h + 1/2(-x2)3/4, > 0, the evidence being that formally it is invariant under the 1:2:3 KPZ scaling (1.1) and preserves Brownian motion. Besides the non-physical non-locality, and the inherent difficulty of making sense of this equation, one can see that it is not correct because it has two free parameters instead of one. Presumably, it converges to the KPZ fixed point in the limit 0. On the other hand, the model has critical scaling, so it is also plausible that if one introduces a cutoff (say, smooth the noise) and then take a limit, the result has = 0, and possibly even a renormalized . So it is possible that, in a rather uninformative sense, the conjecture could still be true. 8This fixes our study of the scaling limit to perturbations of density 1/2. We could perturb off any density (0, 1) by observing in an appropriate moving frame without extra difficulty, but we do not pursue it here. THE KPZ FIXED POINT 11 The left hand side is also taken to be the linear interpolation to make it a continuous function of x R. For fixed t > 0, we will prove that the limit h(t, x; h0) = lim h(t, x) 0 (3.4) exists, and take it as our definition of the KPZ fixed point h(t, x; h0). We will often omit h0 from the notation when it is clear from the context. 3.1. State space and topology. The state space in which we will always work, and where (3.2), (3.3) will be assumed to hold and (3.4) will be proved, in distribution, will be9 UC = upper semicontinuous fns. h : R [-, ) with h(x) C(1 + |x|) for some C < with the topology of local UC convergence, which is the natural topology for lateral growth. We describe this topology next. Recall h is upper semicontinuous (UC) iff its hypograph hypo(h) = {(x, y) : y h(x)} is closed in [-, ) × R. [-, ) will have the distance function10 d[-,)(y1, y2) = |ey1 - ey2|. On closed subsets of R × [-, ) we have the Hausdorff distance d(C1, C2) = inf{ > 0 : C1 B(C2) and C2 B(C1)} where B(C) = xC B(x), B(x) being the ball of radius around x. For UC functions h1, h2 and M = 1, 2, . . ., we take dM (h1, h2) = d(hypo(h1)M , hypo(h2)M ) where M = [-M, M ] × [-, ). We say h - h if h(x) C(1 + |x|) for a C independent of and dM (h, h) 0 for each M 1. We will also use LC = g : -g UC (made of lower semicontinuous functions), the distance now being defined in terms of epigraphs, epi(g) = {(x, y) : y g(x)}. 3.2. For any h0 UC, we can find initial data X0 so that (3.3) holds in the UC topology. This is easy to see, because any h0 UC is the limit of functions which are finite at finitely many points, and - otherwise. In turn, such functions can be approximated by initial data X0 where the particles are densely packed in blocks. Our goal is to take such a sequence of initial data X0 and compute Ph0(h(t, xi) ai, i = 1, . . . , M ) which, from (2.1) and (3.4), is the limit as 0 of PX0 X-3/2t( 1 4 -3/2t - -1xi - 1 2 -1/2ai + 1) 2-1xi - 1, i = 1, . . . , M . (3.5) We therefore want to consider Theorem 2.6 with t = -3/2t, ni = 1 4 -3/2t - -1xi - 1 2 -1/2 ai + 1. (3.6) While (2.26) is more general, it turns out (2.25) is nicer for passing to limits. There is no loss of generality because of the next lemma, which says that we can safely cut off our data far to the right. For each integer L, the cutoff data is X0,L(n) = X0(n) if n > - -1L and X0,L(n) = if n - -1L . This corresponds to replacing h0(x) by h0,L(x) with a straight line with slope -2-1/2 to the right of X0(- -1L ) 2L. The following will be proved in Appendix B.5: Lemma 3.2. (Finite propagation speed) Suppose that X0 satisfies (3.3). There are 0 > 0 and C < and c > 0 independent of (0, 0) such that the difference of (3.5) computed with initial data X0 and with initial data X0,L is bounded by C(e-cL3 1Lc-1/2 + L-1/21L>c-1/2 ). 9The bound h(x) C(1 + |x|) is not as general as possible, but it is needed for finite propagation speed (see Lemma 3.2). With work, one could extend the class to h(x) C(1 + |x|), < 2. Once the initial data has parabolic growth there is infinite speed of propagation and finite time blowup. 10This allows continuity at time 0 for initial data which takes values -, such as half-flat (see Section 4.4). THE KPZ FIXED POINT 12 3.3. The limits are stated in terms of an (almost) group of operators St,x = exp{x2 - t 6 3}, x, t R2 \ {x < 0, t = 0}, (3.7) satisfying Ss,xSt,y = Ss+t,x+y as long as all subscripts avoid {x < 0, t = 0}. We can think of them as unbounded operators with domain C0(R). It is somewhat surprising that they even make sense for x < 0, t = 0, but it is just an elementary consequence of the following explicit kernel and basic properties of the Airy function11 Ai(z) = 1 2i dw e 1 3 w3 -zw . The St,x act by convolution St,xf (z) = - dy St,x(z, y)f (y) = - dy St,x(z - y)f (y) where, for t > 0, 1 St,x(z) = 2i dw e t 6 w3 +xw2 -zw = (t/2)-1/3 e 2x3 3(t/2)2 + 2zx t Ai((t/2)-1/3z + (t/2)-4/3x2), (3.8) and S-t,x = (St,x), or S-t,x(z, y) = S-t,x(z - y) = St,x(y - z). Since |Ai(z)| C e- 2 3 z3/2 for z 0 and |Ai(z)| C for z < 0, St,x is actually a bounded operator on L2(R, dz) whenever x > 0, t = 0. For x 0 it is unbounded, with domain Dx+ = {f L2(R) : 0 dz e2z|x/t|f (z) < } if t > 0 and Dx- = {f L2(R) : 0 - dz e-2z|x/t|f (z) < } if t < 0. Our kernels will always be used with conjugations which put us in these spaces. In addition to St,x we need to introduce the limiting version of S¯te,pni(X0). For g LC, S¯etp,xi(g)(v, u) = EB(0)=v St,x- (B( ), u)1 < , (3.9) where B(x) is a Brownian motion with diffusion coefficient 2 and is the hitting time of the epigraph of g12,13. Note that, trivially, S¯etp,xi(g)(v, u) = St,x(v, u) for v g(0). If h UC, there is a similar operator S¯ht,yxpo(h) with the same definition, except that now is the hitting time of the hypograph of h and S¯ht,yxpo(h)(v, u) = St,x(v, u) for v h(0). Lemma 3.3. Under the scaling (3.6) and assuming that (3.3) holds in LC, if we set zi = 2-1xi + -1/2(ui + ai), y = -1/2v, then we have, as 0, St,xi (v, ui) := -1/2St,-ni (y , zi) - St,xi (v, ui), St,-xj (v, uj ) := -1/2St,nj (y , zj ) - St,-xj (v, uj ), S¯t,,-epxi(j-h-0 )(v, uj ) := -1/2S¯te,pnij(X0)(y , zj ) - S¯etp,-i(x-jh-0 )(v, uj ) pointwise, where h-0 (x) = h0(-x) for x 0. (3.10) (3.11) (3.12) The pointwise convergence is of course not enough for our purposes, but will be suitably upgraded to Hilbert-Schmidt convergence, after an appropriate conjugation, in Lemmas B.4 and B.5. Sketch of the proof of Lemma 3.3. We only sketch the argument, since the results in Appendix B.3 are stronger. We use the method of steepest descent14. From (2.21), where 1 St,-ni (zi, y) = 2i e dw, -3/2F (3)+-1F (2)+-1/2F (1)+F (0) 0 (3.13) F (3) = t (w - 1 2 ) + 1 4 log( 1-w w ) , F (2) = -xi log 4w(1 - w), F (1) = (ui - v - 1 2 ai) log 2w - 1 2 ai log 2(1 - w), F (0) = - log 1-w 2w . (3.14) 11Here is the Airy contour; the positively oriented contour going from e-i/3 to ei/3 through 0. 12It is important that we use B( ) in (3.9) and not g( ) which, for discontinuous initial data, could be strictly smaller. 13St,x-y(B(y), u) is a martingale in y 0. However, one cannot apply the optional stopping theorem to conclude that EB(0)=v St,x- (B( ), u)1< = St,x(v, u). For example, if g 0, one can compute EB(0)=v St,x- (B( ), u)1< = St,x(-v, u). The minus sign is not a mistake! 14We note that this (or rather Appendix B) is the only place in the paper where steepest descent is used. THE KPZ FIXED POINT 13 The leading term has a double critical point at w = 1/2, so we introduce the change of variables w 1 2 (1 - 1/2w~), which leads to -3/2F (3) t 6 w~3 , -1F (2) xiw~2, -1/2F (1) -(ui - v)w~. (3.15) We also have F (0) log(2), which cancels the prefactor 1/2 coming from the change of variables. In view of (3.8), this gives (3.10). The proof of (3.11) is the same, using (2.22). Now define the scaled walk B(x) = 1/2 B-1x + 2-1x - 1 for x Z+, interpolated linearly in between, and let be the hitting time by B of epi(-h(0, ·)-). By Donsker's invariance principle [Bil99], B converges locally uniformly in distribution to a Brownian motion B(x) with diffusion coefficient 2, and therefore (using (3.3)) the hitting time converges to as well. Thus one can see that (3.12) should hold; a detailed proof is in Lemmas B.1 and B.5. We will compute next the limit of (3.5) using (2.24) under the scaling (3.6). To this end we change variables in the kernel as in Lemma 3.3, so that for zi = 2-1xi + -1/2(ui + ai) we need to compute the limit of -1/2 ¯2-1xKt¯2-1x (zi, zj). Note that the change of variables turns ¯2-1x(z) into ¯-a(u). We have ni < nj for small if and only if xj < xi and in this case we have, under our scaling, -1/2Qnj-ni (zi, zj ) - e(xi-xj)2 (ui, uj ), as 0. For the second term in (2.25) we have -1/2(St,-ni )X0(1)St,nj (zi, zj ) = -1 -1/2X0(1) dv (St,xi )(ui, -1/2v)St,-xj (-1/2v, uj ) - (St,xi )-h0(0)St,-xj (ui, uj ). The limit of the third term in (2.25) is proved similarly. Thus we obtain a limiting kernel - e(xi-xj)2 (ui, uj )1xi>xj + (St,xi )-h0(0)St,-xj (ui, uj ) + (St,xi )¯-h0(0)S¯etp,-i(x-jh-0 )(ui, uj ), (3.16) surrounded by projections ¯-a. Our computations here only give pointwise convergence of the kernels, but they will be upgraded to trace class convergence in Appendix B, which thus yields convergence of the Fredholm determinants. We prefer the projections ¯-a surrounding (3.16) to read a, so we change variables ui - -ui and replace the Fredholm determinant of the kernel by that of its adjoint to get det I - aKhexytpo(h0)a with Khexytpo(h0)(ui, uj) = the kernel in (3.16), evaluated at (-uj, -ui) and with xi and xj flipped. But St,x(-u, -v) = (St,x)(v, u), so (St,xj )-h0(0)St,-xi (-uj , -ui) = (St,-xi )¯h0(0)St,xj (ui, uj ). Similarly, we have S¯etp,xi(-h-0 )(-v, -u) = (S¯ht,yxpo(h-0 ))(u, v) for v -h0(0), and thus we get (St,xj )¯-h0(0)S¯etp,-i(x-ih-0 )(-uj , -ui) = (S¯ht,y-pxoi(h-0 ))h0(0)St,xj (ui, uj ). This gives the following preliminary (one-sided) fixed point formula. Theorem 3.4. (One-sided fixed point formulas) Let h0 UC with h0(x) = - for x > 0. Given x1 < x2 < · · · < xM and a1, . . . , aM R, Ph0(h(t, x1) a1, . . . , h(t, xM ) aM ) = det I - aKhexytpo(h0)a L2({x1,...,xM }×R) = det I - Kht,yxpMo(h0) + Kht,yxpMo(h0)e(x1-xM )2 ¯a1 e(x2-x1)2 ¯a2 · · · e(xM -xM-1)2 ¯aM (3.17) L2(R) (3.18) with Khexytpo(h0)(xi, ·; xj , ·) = -e(xj-xi)2 1xiL, which can be obtained from the previous theorem by translation invariance. We then take a continuum limit of the operator e(x1-xM )2 ¯a1 e(x2-x1)2 ¯a2 · · · e(xM -xM-1)2 ¯aM on the right side of (3.18) to obtain a "hit" operator for the final data as well. From Lemma 3.2, the result is the same as if we started with two-sided data for TASEP. The shift invariance of TASEP tells us that h(t, x; hL0 ) d=ist h(t, x - L; LhL0 ), where L is the shift operator from (2.27). Our goal then is to take L in the formula given in Theorem 3.4 for h(t, x - L; LhL0 ). The left hand side of (3.17) becomes det I - aKLLhL0 a with L2({x1,...,xM }×R) KLLhL0 (xi, ·; xj, ·) given by e(xj -xi)2 1xi