* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Download ABSTRACT ACCELERATION AND OBSERVER DEPENDENCE OF
Schrödinger equation wikipedia , lookup
Particle in a box wikipedia , lookup
Double-slit experiment wikipedia , lookup
Renormalization group wikipedia , lookup
Density matrix wikipedia , lookup
Coherent states wikipedia , lookup
Molecular Hamiltonian wikipedia , lookup
Path integral formulation wikipedia , lookup
History of quantum field theory wikipedia , lookup
Dirac equation wikipedia , lookup
Wave–particle duality wikipedia , lookup
Matter wave wikipedia , lookup
Symmetry in quantum mechanics wikipedia , lookup
Scalar field theory wikipedia , lookup
Canonical quantization wikipedia , lookup
Theoretical and experimental justification for the Schrödinger equation wikipedia , lookup
ABSTRACT ACCELERATION AND OBSERVER DEPENDENCE OF VACUUM STATES IN QUANTUM FIELD THEORY We derive a novel flat spacetime metric of a non-uniformly accelerating observer (NUAO). Like the UAOs, these particular NUAOs see a Bose-Einstein distribution. However, we show that these NUAOs, unlike UAOs, experience no event horizons, no temperature, and no information loss. This has implications for black hole evaporation, firewalls, and the information paradox. Alaric Doria May 2016 ACCELERATION AND OBSERVER DEPENDENCE OF VACUUM STATES IN QUANTUM FIELD THEORY by Alaric Doria A thesis submitted in partial fulfillment of the requirements for the degree of Master of Science in Physics in the College of Science and Mathematics California State University, Fresno May 2016 APPROVED For the Department of Physics: We, the undersigned, certify that the thesis of the following student meets the required standards of scholarship, format, and style of the university and the student’s graduate degree program for the awarding of the master’s degree. Alaric Doria Thesis Author Gerardo Muñoz (Chair) Physics Douglas Singleton Physics Karl Runde Physics For the University Graduate Committee: Dean, Division of Graduate Studies AUTHORIZATION FOR REPRODUCTION OF MASTER’S THESIS I grant permission for the reproduction of this thesis in part or in its entirety without further authorization from me, on the condition that the person or agency requesting reproduction absorbs the cost and provides proper acknowledgment of authorship. X Permission to reproduce this thesis in part or in its entirety must be obtained from me. Signature of thesis author: ACKNOWLEDGMENTS I would like to thank my advisor and committee members, Professor Gerardo Muñoz, Professor Douglas Singleton, and Professor Karl Runde, for their helpful advice, guidance and knowledge. TABLE OF CONTENTS Page LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi INTRODUCTION AND BACKGROUND . . . . . . . . . . . . . . . . . 1 Special Relativity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 ACCELERATION IN SPECIAL RELATIVITY . . . . . . . . . . . . . . 6 Derivation of Metric . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 Kinematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 Minkowski Coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . 10 Light Cone Coordinates . . . . . . . . . . . . . . . . . . . . . . . . . 12 UNRUH RADIATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 Quantum Field Theory in the Uniformly Accelerated Frame . . . . . 14 Unruh Radiation Particle Eigenstates Expansion . . . . . . . . . . . . 21 QUANTUM FIELDS AND NON-UNIFORMLY ACCELERATING OBSERVERS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 Quantum Fields and Non-uniformly Accelerating Observers . . . . . . 29 Expansion of the Minkowski Vacuum State . . . . . . . . . . . . . . . 34 CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 FIGURE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 APPENDICES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 Appendix A: Rindler Observers and Unruh Temperature . . . . . . . 42 Appendix B: Verification of Bose-Einstein Distribution of Particle Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 LIST OF FIGURES Page Figure 1. The spacetime diagram for a non-uniformly accelerating observer with coordinates (T, X) plotted with the corresponding Minkowski coordinates (t, x) light cone. The hyperbolae cross the light cone making explicit the absence of horizons. . . . . . . . . . 41 INTRODUCTION AND BACKGROUND A complete description of accelerating observers in flat spacetime is lacking in the literature. Observers moving with constant proper acceleration (Rindler observers [1]) are the most widely used description of such behavior so far, especially in discussions of the thermal properties of vacuum states in quantum field theories [2]. In this work we derive the trajectory and the metric of an observer moving with non-constant proper acceleration in two-dimensional flat spacetime. It is clear that Rindler observers are unphysical. They accelerate with constant proper acceleration for all time, reaching speeds arbitrarily close to the speed of light. For this reason, Rindler observers see the light cone as an event horizon. These horizons partition the spacetime into four regions [1]: the left and right Rindler wedges, which we characterize by a metric ds2 = X 2 dT 2 − dX 2 , and the Kasner-Milne [3, 4, 5] universe, which we characterize by a metric ds2 = dT 2 − T 2 dX 2 . These regions are bounded by the lines in flat spacetime at 45 degrees to the x- and t-axes (in units of c = 1). Observers in the Rindler wedges are disconnected from some of the other wedges and thus do not have complete access to the full spacetime. For instance, an observer in the right Rindler wedge can receive signals from the past of the Kasner-Milne universe (bottom wedge) and send signals to the future Kasner-Milne universe (top wedge), but is completely disconnected from the left Rindler wedge. This is because observers in the Rindler wedges see event horizons corresponding to the light cone portions in their quadrants 2 of the spacetime. Minkowski (inertial) observers never experience this; an event occuring in flat spacetime is always accessible to the Minkowski observer provided the observer waits sufficiently long times. The Kasner-Milne universe is a constantly expanding universe where all points in spacetime are moving at different constant rates, with the exception of the center of the universe. An observer in this spacetime would expand outward, away from the central point of the universe, traveling with constant velocity. The observer’s speed is determined by their initial starting distance from the center of the universe. We are particularly interested in investigating the relationship between vacuum states in quantum field theory and non-inertial observers and therefore focus our attention towards metrics with non-zero acceleration. We begin with a brief introduction to special relativity. We write an invariant measure called the spacetime interval ds consistent with the two postulates of special relativity. As we show, the spacetime interval is fundamental in defining all of the physical quantities in special relativity, such as the proper time. The special relativistic gamma factor γ, which is responsible for all deviations from Newtonian physics, follows naturally from the definition of ds. We use the concept of the proper time to define 4-vectors like the 4-velocity, and the 4-acceleration in analogy with Newtonian physics. Invariant scalars are constructed by taking 4-vector products defined through the metric tensor; an object whose sign signature alters the standard scalar product of two vectors. Finally, we define the invariant proper acceleration. This is the acceleration that is measured the same in all inertial reference 3 frames, that is, those whose proper acceleration is zero. Therefore, the proper acceleration categorically describes the motion of a particle. In the next section, we derive a metric which describes non-uniformly accelerating observers with proper acceleration that decays to zero for long times. Special cases of this metric include both the Kasner-Milne metric and the Rindler metric. We derive the velocity, acceleration, proper acceleration, and trajectory of the observers. We make explicit the fact that these observers accelerate to terminal speeds less than the speed of light. We show analytically, and graphically in Figure 1, that the hyperbolic trajectories these observers experience cross the 45-degree lines representing the Minkowski light cone; the same light cone which Rindler observers see as an event horizon. We note the corresponding light cone coordinates which we use later. Special Relativity Everything in Special Relativity begins with two postulates: The first postulate is that space and time are homogeneous and isotropic. The second is that the Universe has a maximum speed limit; the speed of light. This allows one to define an invariant spacetime interval ds. ds2 = c2 dt2 − dx2 − dy 2 − dz 2 (1) Invariant here means that the spacetime interval between any two events ds takes the same value no matter which frame of reference measures it. The dimensions of the spacetime interval are length; however, defining 4 an invariant time measure, the proper time, as ds = cdτ is natural. Isolating the time dimension on the right hand side yields " 2 # ~v c2 dτ 2 = 1 − c2 dt2 c (2) where the differentials have been replaced by the velocity ~v . We see that the time measured by an inertial observer is related to the proper time by γdτ = dt. Where γ is defined as 2 #− 12 ~v γ ≡ 1− c " (3) and takes values in the range [1, ∞). From this point on, we take the speed of light equal to unity (c = 1). The spacetime interval can be expressed equivalently in terms of the 4-vector displacement dxµ = (dt, d~x) when a metric ηµν = diag (1, −1, −1, −1) is introduced. ds2 = ηµν dxµ dxν = dxν dxν (4) Where we have defined dxν = ηµν dxµ , a 4-vector with a lowered index. Now we define the 4-velocity as the derivative of the 4-displacement with respect to the proper time; this guarantees that it is a 4-vector itself. dxµ ≡ uµ = γ (1, ~v ) dτ (5) There is a corresponding condition on the 4-velocity uµ like in the case of dxµ . 5 The Lorentz invariant for uµ is then uµ uµ = 1 (6) which confirms the second postulate that the speed of light is the maximum speed and is invariant. From the 4-velocity we define the 4-acceleration as the derivative of the 4-velocity with respect to the proper time. This leads to duµ ≡ aµ = γ γ 3 (~v · ~a) , γ 3 (~v · ~a) ~v + γ~a dτ (7) where the γ factor multiplying the components of uµ has made the derivative non-trivial. The vector ~a = d~v /dt represents the 3-acceleration. Again, we construct an invariant from the new 4-vector aµ aµ = a2pr = γ 6 a2 where we have labeled the invariant apr meaning proper acceleration. Therefore, the proper acceleration, being an invariant, characterizes the accelerating motion for all external observers. (8) ACCELERATION IN SPECIAL RELATIVITY Without loss of generality it suffices to work with a two dimensional metric; this is because the accelerating observers under consideration accelerate in a fixed direction which can always be oriented to the x-direction using a coordinate transformation. Derivation of Metric We begin with the flat spacetime metric ds2 = dt2 − dx2 (9) and change coordinates to the accelerating frame. The most general coordinate transformation is t = h(ξ, ζ) and x = f (ξ, ζ). These coordinate transformations lead to the following metric: ds2 = (ḣ2 − f˙2 )dξ 2 + 2(ḣh0 − f˙f 0 )dξdζ + (h02 − f 02 )dζ 2 (10) Here primes indicate derivatives with respect to ζ and dots indicate derivatives with respect to ξ. We attempt to eliminate the cross term dξdζ and consider a metric with a time-like and space-like coordinate; later, we consider a case where the cross term survives which is likened to the light cone coordinates. We choose ḣh0 = f˙f 0 and solve the partial differential equation. We assume a solution that is separable, t = a(ξ)b(ζ) and 7 x = c(ξ)d(ζ). It follows that ȧa dd0 = 0 ċc bb (11) Each side of the equation depends only on ξ or ζ and therefore must be equal to a constant. Integrating both equations, we obtain relationships between the functions a and c as well as d and b. a2 − k 2 c2 = A2 (12) d 2 − k 2 b 2 = B 2 The general solutions to the equations in (12) come in four cases. We consider the trivial case first, A = B = 0. This leads to functions a = ±kc and d = ±kb and forces ds2 = 0. We take cases two and three to be A2 = 0 and B 2 6= 0, and vice versa. First we take A2 = 0; this reduces the equations in (12) to a = ±kc and (d/B)2 − (kb/B)2 = 1. The general solution to this equation is d = B cosh φ(ζ) and b = Bk −1 sinh φ(ζ). Thus, our coordinate transformations look like t = Bc(ξ) sinh φ(ζ) and x = Bc(ξ) cosh φ(ζ), and we see that case two corresponds to the left and right Rindler wedges. With coordinate choices X = Bc(ξ) and T = φ(ζ) the metric becomes ds2 = X 2 dT 2 − dX 2 (13) For case three, the condition A2 6= 0 and B 2 = 0 implies t = Ab(ζ) cosh θ(ξ) and x = Ab(ζ) sinh θ(ξ), which describe the Kasner-Milne 8 universe. The coordinate choices X = θ(ξ) and T = Ab(ζ) now yield the metric ds2 = dT 2 − T 2 dX 2 (14) Observers in the Kasner-Milne wedges move with constant velocity; the distance between two observers decreases in the lower wedge and increases in the upper wedge, leading some to refer to the upper and lower wedges as the expanding and contracting universe, respectively [2]. The fourth – and most interesting – case is A2 6= 0 and B 2 6= 0. Then, the solutions to equations (12) with the condition that A and B are both non-zero become a = A cosh φ(ξ), c = Ak −1 sinh φ(ξ), and also d = B cosh θ(ζ), b = Bk −1 sinh θ(ζ). With the appropriate choice of X and T it follows that t = α cosh X α T sinh α , x = α sinh X α T cosh α (15) where α = ABk −1 . The metric of the accelerating observer is therefore 2 ds = cosh X +T α cosh X −T α (dT 2 − dX 2 ) (16) In Figure 1 we show a spacetime diagram with lines of constant T and constant X. It is clear from this figure that the accelerating observer sees no event horizons. The trajectories are hyperbolic but they are not bounded by the light cones. 9 Kinematics We select a fixed position X0 in the instantaneous rest frame of the accelerating observer and calculate the velocity. dx = tanh dt X0 α T T tanh = v∞ tanh α α (17) The speed of the accelerating observer ranges from 0 to 1, the speed of light, and we may adjust this speed by varying either the position of the accelerating observer in his coordinate system or the parameter α. We define the terminal velocity v∞ ≡ tanh(X0 /α) for convenience and physical insight, and denote by γ∞ the relativistic gamma factor associated with this terminal velocity, so that γ∞ = cosh(X0 /α). We then calculate the acceleration a = d2 x/dt2 . a= v∞ α cosh 1 X0 cosh3 α = T α v∞ 1 αγ∞ cosh3 T α (18) The transformation law to obtain the proper acceleration is apr = γ 3 a. apr −3/2 v∞ T 2 −2 = 1 + γ∞ sinh αγ∞ α (19) For completeness, we note from (16) or from the equivalent form T X 2 2 ds = cosh + sinh (dT 2 − dX 2 ) α α 2 that the relationship between the time T and the proper time τ of the (20) 10 observer at X0 is −1/2 T 2 2 dT = γ∞ + sinh dτ α (21) Unlike in the Rindler case, dT and dτ are related here by a time-dependent factor. We have yet to identify the meaning of the parameter α. To determine what α is physically, we need to analyze the trajectory x(T ) which we obtain from the original coordinate transformations (15). x(T ) = α sinh X0 α 1/2 T 2 1 + sinh α (22) We also see from (15) that t = 0 ⇒ T = 0, and this is the exact condition we need to determine α. x0 = α sinh X0 α = αγ∞ v∞ (23) Therefore, α = x0 /v∞ γ∞ is determined by the distance of closest approach x0 and the initial velocity v∞ . Alternatively, we could also fix α by specifying the acceleration at T = t = 0 which leads to α = v∞ /γ∞ a0 , where a0 = a(0) = apr (0). In contrast to Rindler, where observers require infinite accelerations [1], X0 = 0 corresponds to stationary observers. Minkowski Coordinates We now return to our original kinematic equations and replace α with physical parameters and restore Minkowski time. This allows us to make comparisons with what we know about the equations of motion for a 11 uniformly accelerating observer. The trajectory becomes " x(t) = x0 1 + v∞ t x0 2 #1/2 (24) For large t, we see that (24) reduces to the equation of a line x(t) = v∞ t with slope v∞ . This explicitly shows that the hyperbolas cross over the 45-degree lines that define event horizons for uniformly accelerating observers. The same conclusion follows from the velocity " 2 #−1/2 2 v∞ t t v∞ 1+ v(t) = x0 x0 (25) by considering again large t behavior, where v becomes v∞ . The acceleration is deceptively similar to that of a uniformly accelerating observer. " 2 #−3/2 2 v∞ t v∞ 1+ a(t) = x0 x0 (26) We recover the exact answer for the acceleration of a uniformly accelerating observer by taking the limit as v∞ → 1 with x0 (equivalently, αγ∞ ) fixed. The proper acceleration is " " 2 #−3/2 2 #−3/2 2 v∞ v∞ t v∞ t apr (t) = 1+ = 1+ 2 x0 x0 γ∞ αγ∞ αγ∞ (27) which clearly describes a non-uniformly accelerating observer. We see exactly what happens in the case of uniform acceleration in equation (27) by taking 12 the same limit v∞ → 1 with x0 fixed. The proper acceleration reduces to the constant apr = 1/x0 one would expect for a uniformly accelerating observer because the γ∞ dividing the time coordinate becomes infinite as we take the limit. The transformation apr = γ 3 a from acceleration to proper acceleration is equivalent to dividing the time coordinate t by a single power of γ∞ due to the fact that the gamma factor resulting from (25) is given by " 1 + (v∞ t/x0 )2 γ(t) = 1 + (v∞ t/x0 γ∞ )2 #1/2 (28) where we recover the exact γ factor for the uniformly accelerating observer in the limit v∞ → 1 as 1/2 γ(t) = 1 + (t/x0 )2 (29) where the entire denominator of (28) becomes one due to γ∞ → ∞. Light Cone Coordinates We denote by u = t − x and v = t + x the standard light cone coordinates in Minkowski spacetime. We define U = T − X and V = T + X and use hyperbolic identities to obtain the following. U u = α sinh α , V v = α sinh α (30) These coordinates are completely invertible since the hyperbolic sine inverse is a well defined function on the entire domain. This isolates the coordinates, leaving u as a function of U only, and v as a function of V only. Note that 13 u = 0 ⇐⇒ U = 0 and v = 0 ⇐⇒ V = 0, and that both coordinate systems cover the entire spacetime. In light cone coordinates the metric becomes U V ds = cosh cosh dU dV α α 2 (31) These results will be crucial in determining whether non-uniformly accelerating observers can associate thermal properties with the Minkowski vacuum [12]. UNRUH RADIATION Introduction Soon after the discovery, by Bekenstein and Hawking [6, 7], that black holes possess entropy and temperature, Davies [8] and Unruh [9], following the work of Fulling [10], showed that observers moving with constant proper acceleration in flat spacetime see the Minkowski vacuum as a thermal bath with temperature T = ~a/2πk [8, 9]. The constants ~ and k are Planck’s constant and Boltzmann’s constant respectively. The Unruh effect has deservedly become the prime flat-spacetime example illustrating a deep connection between relativity, quantum mechanics, and thermodynamics. Bekenstein postulated that the entropy of matter falling into a black hole should contribute entropy to the black hole itself [7]. This eventually led to the proportionality of the entropy S of a black hole to the area of it’s event horizon A [6, 7]. Hawking showed that the proportionality went like S = A/4 and that the temperature associated with his entropy was T = ~κ/2πk where κ is the surface gravity of the black hole. Unruh showed that the same effect applied to Rindler observers in a flat spacetime background. Quantum Field Theory in the Uniformly Accelerated Frame A uniformly accelerating observer in special relativity is characterized by the Rindler observer. This observer accelerates arbitrarily close to the speed of light and sees the light cone as an event horizon. Due to the presence of the horizons in the accelerating frame the spacetime for the Rindler 15 observer is partitioned into four regions; the two regions, left and right, characterize the uniformly accelerating observer, while the top and bottom wedges characterize an observer with zero acceleration. We will focus on the left and right regions where the observer is accelerating. Due to the partitioning of the spacetime, the accelerating observer sees radiation in the Minkowski vacuum state. The radiation has temperature T = ~α/2πk. As discussed previously, we describe the Rindler observer by the following metric. ds2 = X 02 dT 02 − dX 02 (32) We label the spatial coordinate X 0 for now, and reserve the label X for the final spatial transformation; the same is done for T 0 . The coordinates X 0 and T 0 are described in terms of the inertial x and t as follows: t = X 0 sinh T 0 , x = X 0 cosh T 0 (33) We calculate the velocity v = tanh(T 0 ) which tends arbitrarily close to one. Therefore, this observer sees the light cone as an event horizon. The proper acceleration is then apr = x−1 0 = α where x0 is the point on the x-axis which intersects t = 0. Uniformly accelerating observers are then solely described by the parameter x0 which characterizes the magnitude of the proper acceleration. It will be most convenient to solve for the quantum field operator in light cone coordinates because we will be considering a massless spin zero field which corresponds to an operator differential equation like the wave equation 16 for functions. The wave operator, which generates the wave equation when acting on functions (or operators) is invariant under conformal transformations; this allows for an easily expressible general solution for the quantum field operator. In order to put the metric into a conformal form, we introduce two additional coordinate transformations: T 0 = αT and X 0 = α−1 exp(αX). ds2 = e2αX dT 2 − dX 2 (34) However, the transformation X 0 = α−1 exp(αX) forces the X 0 coordinate to be positive. Therefore, an additional case for X 0 = −α−1 exp(αX) is required and addressed later when necessary. Now we consider a massless Klein-Gordon scalar quantum field theory. As previously described, the differential equation for the quantum field operator is reduced to the wave equation for an operator. In the (X, T ) coordinates, the wave equation is ∂ 2 Φ̂ ∂ 2 Φ̂ − =0 ∂T 2 ∂X 2 (35) where we have divided out the conformal exponential factor because it is always greater than zero. Notice that the field Φ̂ is a scalar field operator, not a scalar function. Therefore, the physical quantum fields that are observed by experiments are expectation values of the operator Φ̂. We write the solutions for Φ̂ in terms of the light cone coordinates U = T − X and V = T + X. Φ̂(U, V ) = Φ̂1 (U ) + Φ̂2 (V ) (36) 17 From here, we consider only the left moving modes Φ̂2 (V ) as the analysis for the right moving modes is essentially identical. We relabel Φ̂2 (V ) = Φ̂(V ) as we will only be working with Φ̂2 (V ). The solution can be written in terms of the frequency modes given by the Fourier expansion as ∞ Z Φ̂(V ) = 0 i dΩ h R −iΩ(T +X) R† iΩ(T +X) √ b̂Ω e + b̂Ω e 4πΩ (37) R† The symbols b̂R Ω and b̂Ω are called the creation and annihilation operators. i h R† , b̂ They satisfy the discretized commutator relationship b̂R Ω Ω0 = δΩΩ0 where δΩΩ0 is the Kronecker delta and is one when the frequencies Ω and Ω0 are the same, and zero otherwise. These operators are labeled by the frequency Ω which describes the frequency of the particles that they “create” or “annihilate” when acting on particle eigenstates. The superscript R indicates the region of the spacetime in which the operator acts; in this case, the operator acts on particle eigenstates only in the right Rindler wedge. The R† action of b̂R Ω and b̂Ω on particle eigenstates is of utmost importance when discussing observer dependence of vacuum states and will be discussed in more detail later in the thesis. The case when X 0 = −α−1 exp(αX) will lead to a similar result; however, the sign of X in equation (37) changes to a minus sign. This yields the solution for the left wedge of the spacetime. Z Φ̂(V ) = 0 ∞ i dΩ h L −iΩ(T −X) iΩ(T −X) √ b̂Ω e + b̂L† e Ω 4πΩ Notice that the label on the creation and annihilation operators is now L (38) 18 meaning that these operators act on particle eigenstates only in the left Rindler wedge. They satisfy a similar commutation relationship h i L L† b̂Ω , b̂Ω0 = δΩΩ0 . All of the commutators are equal to zero if the superscripts h i R L† R and L do not match, b̂Ω , b̂Ω0 = 0, because the operators act on particle eigenstates in different parts of the Rindler spacetime. Since the quantum field operator Φ̂ is a scalar operator, a similar solution can be written in terms of the inertial coordinates and these solutions must yield equality. Z 0 ∞ Z ∞ i dω dΩ h R −iΩV iΩV −iωv † iωv √ √ b̂Ω e + b̂R† e âω e + âω e = θ(V ) Ω 4πω 4πΩ Z 0∞ i (39) dΩ h L −iΩV 0 iΩV 0 √ +θ(−V ) b̂Ω e + b̂L† e Ω 4πΩ 0 Here, θ(V ) is a Heaviside function which is equal to unity when the argument is positive and zero when the argument is negative. They essentially turn on and off the left and right wedge solutions for Φ̂. We superimpose the two solutions, which are valid in different regions, and obtain a solution which is valid in both regions. The ω on the left hand side of the equation is the frequency of a particle created by the creation and annihilation operators which act on particle eigenstates in the inertial Minkowski spacetime. We will solve for the creation and annihilation operators using orthogonal functions. The following integral formulas are necessary: Z ∞ −∞ 0 e±iωv eiω v dv = 2πδ(ω ± ω 0 ) (40) 19 which is the delta function representation via Fourier transforms, and Z ∞ θ(v)e ±iΩV iωv e −∞ Z dv = ∞ (av)±iΩ/a eiωv dv (41) 0 where we have replaced the coordinate V = a−1 ln(av). The integral in equation (41) has a closed form in terms of Γ functions. To arrive at this closed form we introduce a cutoff in the allowed frequencies ω and analytically continue to the complex plane. The closed form is given by Z ∞ ±iΩ/a iωv (av) e 0 1 ±πΩ/2a ω −iΩ/a Ω dv = e Γ i a a a (42) Combining the results of equations (40) and (42) we isolate the creation and annihilation operators for the Minkowski observer. There are two sets of equations: R∞ −πΩ/a L† b̂R b̂Ω = 1 − e−2πΩ/a 0 dωχâω Ω −e (43) R∞ −πΩ/a L b̂R† b̂Ω = 1 − e−2πΩ/a 0 dωχ∗ â†ω Ω −e and R∞ −2πΩ/a b̂LΩ − e−πΩ/a b̂R† = 1 − e dωχ∗ âω Ω 0 (44) R∞ L† −2πΩ/a b̂Ω − e−πΩ/a b̂R = 1 − e dωχâ†ω Ω 0 where 1 χ= 2πa r Ω πΩ/2a ω −iΩ/a Ω e Γ i ω a a (45) and χ∗ is the complex conjugate of χ. Since âω is the annihilation operator for the Minkowski particle eigenstates âω |0M i = 0 defines the vacuum state for the Minkowski observer. The 0 in the particle eigenstate |0M i means that 20 there are zero particles contained in the state and the M subscript indicates that this particle eigenstate is for the Minkowski observer. The integrals on the right hand side of equations (43) and (44) diverge; however, the Minkowski annihilation operator, when acting on the vacuum state, gives zero and leads to h i −πΩ/a L† b̂R − e b̂ Ω Ω |0M i = 0 i h b̂LΩ − e−πΩ/a b̂R† |0M i = 0 Ω (46) from which it follow that N̂ΩR |0M i = N̂ΩL |0M i (47) R L† L L where N̂ΩR = b̂R† Ω b̂Ω and N̂Ω = b̂Ω b̂Ω are called number operators. We see that there are equal numbers of particles in the left and right particle eigenstates. This means that there is a one-to-one correlation between the particles in the left and right Rindler wedges. The number operator for the Minkowski observer N̂ω returns the particle number as an eigenvalue of the particle eigenstates in the following manner. N̂ω |nM i = n|nM i , n = 0, 1, 2 ... (48) The particle eigenstate |nM i represents a quantum state with n Minkowski particles. An obvious question to ask is ”How many particles does a Rindler observer see in the Minkowski vacuum state?”. We calculate the expectation value h0M |N̂ΩR |0M i using the relationships in equations (46) and recall that 21 the answer is the same for N̂ΩL . −1 h0M |N̂ΩR |0M i = e2πΩ/a − 1 (49) Therefore, a Rindler observer sees a Planckian distribution in the Minkowski vacuum. This is the Unruh radiation. The particles have energy E = ~Ω. Then it is natural to define a temperature by setting the factor 2πΩ/a = E/kT . This leads to the Unruh temperature T = ~a/2πk. In order to see that this definition of temperature is valid, we show that the density matrix is indeed thermal. Unruh Radiation Particle Eigenstates Expansion The Minkowski vacuum state is clearly not a vacuum state for the accelerating observer. The observer dependence of the vacuum state is manifest through the Unruh effect. R We use the notation |{nL , nR }i = |nLΩ1 , nLΩ2 ...i ⊗ |nR Ω1 , nΩ2 ...i for the particle eigenstates of the accelerating observer. We expand the Minkowski vacuum state in terms of the accelerating particle eigenstates. |0M i = X |{nL , nR }ih{nL , nR }|0M i (50) {nL ,nR } Where the sum above runs over all possible combinations of nL and nR corresponding to each particle with frequency Ω. However, recall that the particle numbers in the left and right wedges are the same and therefore nL = nR ≡ n. Consequently, every particle in the right Rindler wedge with 22 frequency Ω has a correlated particle with frequency Ω in the left Rindler wedge. This leads to a simplification in notation that proves to be useful. |0M i = X |{n, n}ih{n, n}|0M i (51) {n} The Rindler particle eigenstates are built from the action of the Rindler L† creation operators b̂Ω and b̂R† Ω on the Rindler vacuum |0R i. Since the number of particles with frequency Ωi in the left and right Rindler wedges are the same, we use b̂†Ωi ≡ b̂†i where i indexes the frequency of the particles being created and nΩi ≡ ni . We express the particle eigenstates for the Rindler observer as |{n, n}i = Y 1 L† ni R† ni b̂i b̂i |0R i (n )! i i (52) where |0R i is the Rindler vacuum state. Therefore, we calculate the inner product h{n, n}|0M i in equation (51) by taking the adjoint of equation (52) and multiplying on the right by the Minkowski vacuum state |0M i. h{n, n}|0M i = ni ni Y 1 b̂R h0R | b̂Li |0M i i (n )! i i (53) From equations (46) we write ni ni ni ni −ni πΩi /a |0 i = e b̂Li b̂L† |0M i b̂Li b̂R M i i (54) where we use the commutator for b̂Li and b̂L† i to write the expression above in 23 terms of Number operators N̂iL . ni ni L† L b̂i b̂i |0M i = ni + N̂i (ni − 1) + N̂i ... 1 + N̂i |0M i (55) Now we rewrite equation (53) using equation (55) where the number operators act to the left on the Rindler vacuum state. This leads to the following expression. h{n, n}|0M i = ∞ YX i e−ni πΩi /a h0R |0M i (56) ni =0 This is substituted into the expression for the Minkowski vacuum given in equation (50). |0M i = ∞ YX i e−ni πΩi /a h0R |0M i |nLi i ⊗ |nR i i (57) ni =0 The normalization of h0M |0M i = 1 and the following summation ∞ X e−2ni πΩi /a = 1 − e−2πΩi /a (58) (59) ni =0 leads to the expression |h0R |0M i|2 = Y i 1 − e−2πΩi /a 24 Finally, the Minkowski vacuum state is ∞ Yp X −2πΩ /a i |0M i = 1−e e−ni πΩi /a |nLi i ⊗ |nR i i (60) ni =0 i By factoring out part of the exponential underneath the square root, we obtain s |0M i = ∞ Ωi X −(ni + 1 )πΩi /a L 2 e 2 sinh π |ni i ⊗ |nR i i a n =0 Y i (61) i This form for the Minkowski vacuum state is useful in making a statistical mechanics interpretation of Unruh radiation. Let Z= Y 1 − e−2πΩi /a −1 (62) i and recall equation (52). This leads to an expression of the Minkowski vacuum state in terms of the Rindler vacuum state. |0M i = Z −1/2 ∞ YX e−ni πΩi /a L† ni R† ni b̂i b̂i |0R i (ni )! i n =0 (63) i The summation is evaluated and we obtain |0M i = Z −1/2 Y R† exp e−πΩi /a b̂L† b̂ |0R i i i (64) i The exponential within the summation of equation (61) can be rewritten to 1 1 e−(ni + 2 )πΩi /a = e−(ni + 2 +ni + 2 )πΩ/2a 1 (65) 25 Thus making explicit the Hamiltonian structure for two independent oscillators; however, the exponent is missing a factor of i. In any case, a prescription for going from a quantum field theory to a statistical mechanics interpretation is to analytically continue β = (kT )−1 = it/~. This analytic continuation associates the statistical mechanical factor exp (−βEi ) to the quantum mechanical evolution operator exp (−iHi t/~) for the harmonic oscillators. Then, from the statistical mechanical point of view, we define a temperature T by associating βEi = 2πΩi a (66) with the appropriate energies Ei = ~Ωi . This leads to the same temperature as before T = ~a/2πk. This is also apparent when we note that equation (59) can be identified as the reciprocal of the partition function Z −1 , where the partition function Z= Y 1 − e−2πΩi /a −1 (67) i is that of a gas of massless particles. This further motivates the association in equation (66). This becomes especially clear in the density matrix formulation. The density of states matrix is ρ = |0M ih0M | (68) which is representative of a pure quantum state; the density matrix is pure, as opposed to thermal. This is also clear when we note ρ2 = ρ, which is 26 characteristic of a pure density matrix. The expectation value of an arbitrary operator A is of course h0M |A|0M i. Consider the case when A is restricted to the right wedge. From equation (64) we write |0M i = Z −1/2 X e− P ni πΩi /a i |{nL }i ⊗ |{nR }i (69) {n} and the expectation value of A is h0M |A|0M i = Z −1 X e− P i (ni +n0i )πΩi /a h{n0L }| h{n0R }|A|{nR }i |{nL }i (70) {n,n0 } where the first summation is over all possible combinations of sets of n and n0 . Since the expectation value is just a constant, the product between the left states gives a Kronecker delta in n and n0 . The summation then reduces to a single summation over all possible combinations of sets of n. This leads to the following. h0M |A|0M i = Z −1 X e− P i ni 2πΩi /a h{nR }|A|{nR }i (71) {n} Here, we associate 2πΩi /a = βEi and bring the exponential inside the product of the states. Since we are working with eigenstates of the Hamiltonian, we write h0M |A|0M i = Z −1 X R h{nR }|e−βH A|{nR }i (72) {n} where H R is the Hamiltonian that acts on the eigenstates in the right wedge. 27 This is exactly the trace of the operators. R tr e−βH A h0M |A|0M i = tr e−βH R (73) The partition function Z has been replaced with an equivalent expression in terms of the trace. This makes explicit the thermal nature of the expectation value of an operator which is restricted to one region of the spacetime. The expectation value is no longer that of a pure quantum state but is instead a thermal average. QUANTUM FIELDS AND NON-UNIFORMLY ACCELERATING OBSERVERS Introduction In this thesis we show that Rindler observers (i.e., observers moving in flat spacetime with constant proper acceleration) are rather special in their identification of the Minkowski vacuum as a thermal state. We make use of the non-uniformly accelerating observer (NUAO) solution derived in equation (16), where observers accelerate to terminal velocities less than the speed of light, to study the properties of the Minkowski vacuum seen by this accelerating observer. We give a brief review of quantum field theory in non-inertial reference frames, followed by a detailed calculation of the creation and annihilation operators seen by our NUAOs. We show that these accelerating observers see a Bose-Einstein distribution of particles, which strongly suggests the definition of a pseudo-temperature in analogy with the Unruh temperature. We justify that this is indeed a pseudo-temperature. We solve for the complete expansion of the Minkowski vacuum in terms of the accelerating observer’s particle number states. Furthermore, we prove that our NUAOs experience the Minkowski vacuum as a single-mode squeezed state and not the typical two-mode squeezed state seen by Rindler observers. This single-mode pure quantum state does not allow for a natural reduction to a mixed state, and the thermal state characteristic of the Unruh effect never enters the physics detected by the NUAO. The absence of a thermal state invalidates the seemingly natural definition of temperature suggested by the 29 Bose-Einstein distribution of particles measured by the NUAO. Finally, a review concerning the differences between Rindler observers and our NUAO is given in the conclusion of the thesis. Additionally, We discuss the physics of our NUAOs in the context of squeezed states, entanglement, black hole physics, and the physical implications thereof. A final remark on general accelerations is given. Quantum Fields and Non-uniformly Accelerating Observers Like the case with constant proper acceleration discussed above, we begin with a massless scalar field and show that non-uniformly accelerating observers see a Bose-Einstein distribution for the expectation value of the number operator hNΩ i in the Minkowski vacuum. ∂ 2 Φ̂ ∂ 2 Φ̂ − =0 ∂t2 ∂x2 (74) We change to light cone coordinates u = t − x and v = t + x. The wave equation becomes ∂ 2 Φ̂ =0 ∂u∂v (75) Φ̂ = Φ̂1 (u) + Φ̂2 (v) (76) The general solution to (75) is As discussed previously, the left (Φ̂1 (v)) and right (Φ̂2 (u)) moving solutions to the field equation are non-interacting; again, we analyze the left-moving solution Φ̂2 (v) = Φ̂(v). We expand the field in terms of Fourier modes as 30 follows: Z Φ̂(v) = 0 ∞ dω √ âω e−iωv + â†ω eiωv 4πω (77) where the creation (â†ω ) and annihilation (âω ) operators for the Minkowski observer obey the standard (discretized) commutation relation [âω , â†ω0 ] = δωω0 . The non-uniformly accelerating observer (NUAO) under consideration sees the following metric: 2 ds = cosh X +T α cosh X −T α (dT 2 − dX 2 ) (78) where the parameter α is related to the acceleration and the initial velocity of the observer. For light cone coordinates U = T − X and V = T + X in the accelerating frame, the metric becomes U V cosh dV dU ds = cosh α α 2 (79) The relationship between (U, V ) and the corresponding Minkowski coordinates (u, v) is given by [11] U u = α sinh α , V v = α sinh α (80) Note that both coordinate systems (u, v) and (U, V ) cover the entire spacetime and there is no singularity located at the origin in the accelerating frame [11]. In other words, the acceleration of the NUAO does not become infinite as we approach the origin. Instead, the acceleration tends to zero. The singularity at the origin for a Rindler observer is responsible for the partitioning of the 31 spacetime and thus the necessity of the left and right creation and annihilation operators. Therefore, in the absence of a singularity, only a single pair of creation and annihilation operators for this NUAO is necessary. The equation satisfied by the scalar field in the accelerating observer’s coordinates U and V is formally identical to (75). ∂ 2 Φ̂ =0 ∂U ∂V (81) The solution is then analogous to (76). Selecting again the left-moving solution we write Z ∞ Φ̂(V ) = 0 i dΩ h √ b̂Ω e−iΩV + b̂†Ω eiΩV 4πΩ (82) in terms of the Fourier modes. The creation (b̂†Ω ) and annihilation (b̂Ω ) operators in the accelerating frame satisfy the same commutation relation [b̂Ω , b̂†Ω0 ] = δΩΩ0 and act on all regions of the spacetime. We identify equations (77) and (82) and solve for the creation and 0 annihilation operators in the accelerating frame. We multiply by e−iΩ V and integrate over V to obtain b̂†Ω . b̂†Ω √ Z ∞ Z ∞ i dω h Ω √ âω e−iωv e−iΩV + â†ω eiωv e−iΩV dV = 2π −∞ ω 0 (83) We refer to equations in (80) and replace v = α sinh(V /α) in the exponentials inside the integral. The integral formula for the modified Bessel function of 32 the second kind [13] 1 Kν (x) = e−iνπ/2 2 Z ∞ dt e−ix sinh t−νt (84) −∞ is valid when x > 0 and −1 < <(ν) < 1. Therefore, using the fact that K−ν = Kν , we have b̂†Ω α√ = Ω π Z ∞ 0 dω √ KiΩα (ωα) e−πΩα/2 âω + eπΩα/2 â†ω ω (85) This integral is divergent and a regularization procedure must be adopted. However, the details do not matter for this calculation, since our results will be independent of the specific regularization procedure and only depend on action of the operators on vacuum states. Denoting by χ(ω) the combination α χ(ω) ≡ π r Ω KiΩα (ωα) ω (86) we write the expression for b̂†Ω more compactly as b̂†Ω =e −πΩα/2 Z ∞ πΩα/2 Z ∞ dωχ(ω) â†ω dωχ(ω) âω + e 0 (87) 0 Similarly, taking the adjoint of equation (87) gives the following expression for b̂Ω . πΩα/2 Z b̂Ω = e ∞ dωχ(ω) âω + e 0 −πΩα/2 Z ∞ dωχ(ω) â†ω (88) 0 We may now isolate the operator âω that annihilates the Minkowski vacuum 33 |0M i by solving the system of equations (87) and (88) for âω . e πΩα b̂Ω − b̂†Ω = e 2πΩα −1 Z ∞ dωχ(ω) âω (89) 0 Therefore, the combination b̂Ω − e−πΩα b̂†Ω annihilates the Minkowski vacuum, b̂Ω − e−πΩα b̂†Ω |0M i = 0 (90) We construct the number operator NΩ = b̂†Ω b̂Ω for the accelerating observer and determine the average particle number h0M |b̂†Ω b̂Ω |0M i. Using equation (90) we find that the norm of the state b̂Ω |0M i is h0M |b̂†Ω b̂Ω |0M i = h0M | e−2πΩα b̂Ω b̂†Ω |0M i (91) Finally, we use the commutator [b̂Ω , b̂†Ω ] = 1 on the right side of equation (91) to obtain −1 h0M |b̂†Ω b̂Ω |0M i = e2πΩα − 1 (92) Equation (92) shows that the non-uniformly accelerating observer sees a Bose-Einstein distribution in the Minkowski vacuum. One would therefore expect to be able to define a temperature from 2πΩα = E/kT and E = ~Ω as T = ~ 2πkα (93) directly from the Bose-Einstein statistics. Although the association of the state with a temperature seems justified, we show that this is not the case. 34 Expansion of the Minkowski Vacuum State We expand the Minkowski vacuum state in terms of the accelerating observer’s number eigenstates. |0M i = X |{ni }ih{ni }|0M i (94) {ni } The states |{ni }i are calculated by repeated action of the b̂†Ω creation operators on the accelerating observer’s vacuum state |0i. We abbreviate b̂Ωi = b̂i and write h{ni }|0M i in terms of products of annihilation operators acting to the left. h{ni }|0M i = √ 1 h0|(b̂1 )n1 (b̂2 )n2 ...|0M i n1 ! · n2 !... (95) Equation (90) in combination with repeated applications of the commutator [b̂Ω , b̂†Ω ] = 1 yields bn |0M i = e−πΩα (N̂ + n − 1)b̂n−2 |0M i (96) where N̂ = b̂† b̂ is the number operator. Iterating the process for b̂n−2 , b̂n−4 , . . . we find h{ni }|0M i = Q j −1)!! h0|0M i j e−nj πΩj α/2 (n√ (nj even) 0 (nj odd) nj ! (97) 35 We relabel ni by 2ki where ki is the number of pairs of particles; this allows us to write a summation over all ki . The expansion of the Minkowski vacuum becomes X Y (2ki − 1)!! −ki πΩi α p e |{2ki }i |0M i = h0|0M i (2ki )! i {ki } (98) We calculate the transition amplitude h0|0M i by left-multiplying by h0M |, replacing h0M |{2ki }i with our result from equation (97) and switching the order of product and summation. 2 1 = h0M |0M i = |h0|0M i| ∞ YX i e−2ki πΩi α ki =0 [(2ki − 1)!!]2 (2ki )! (99) Each sum in equation (99) is given by the closed form ∞ 2 X −1/2 −2πΩi α k [(2k − 1)!!] (e ) = 1 − e−2πΩi α (2k)! k=0 (100) Substitution of the closed form (100) into equation (99) leads to h0|0M i = Y 1 − e−2πΩi α 1/4 ≡ Z −1/2 (101) i where we have dropped an irrelevant phase in extracting the square root. Replacing this result for the vacuum-to-vacuum transition amplitude into equation (98) determines the expansion of the Minkowski vacuum in terms of pairs of bosons to be |0M i = Z −1/2 X Y (2ki − 1)!! −ki πΩi α p |{2ki }i e (2ki )! i {ki } (102) 36 We may now elucidate the relationship between the two vacua. Writing equation (102) as an operator relationship between the two vacuum states is easily accomplished, |0M i = Z −1/2 Y i 1 −πΩi α † 2 exp e bΩi |0i 2 (103) With some effort, one can show that equation (103) is equivalent to |0M i = Y i exp n 1 − ln tanh 4 πΩi α 2 h 2 io 2 † bΩi − bΩi |0i (104) Note that the operator acting on |0i is unitary and is in fact the product of squeezing operators for single-mode states [14]. This shows explicitly that the Minkowski vacuum |0M i is a single-mode squeezed |0i vacuum, and equation (92) may now be seen to be in agreement with a well-known result in quantum optics. CONCLUSIONS We have a metric that describes observers moving with non-constant proper acceleration. Our solution reduces to the Rindler case in the limit v∞ → 1 with x0 fixed. For v∞ 6= 1, the proper acceleration vanishes as |t| → ∞ and is at most the Rindler value apr = 1/x0 . These observers asymptotically approach speeds less than the speed of light, characterized by v∞ , and therefore do not experience event horizons, temperature, or information loss. For Rindler observers, the Minkowski vacuum is perceived as a pure two-mode squeezed state. However, Rindler observers see an event horizon which forces a tracing, in either the left or right wedge, over inaccessible entangled degrees of freedom. The loss of entangled information reduces the pure density matrix ρ = |0M ih0M | to a thermal mixture. This is due to the unphysical nature of constant proper acceleration; such an observer requires an infinite amount of energy to maintain its motion, in stark contrast to the NUAO which requires finite energy. Restricting the constant acceleration to a finite time interval solves the problem of infinite energies, but it does not guarantee a valid definition of temperature. As a matter of fact, not even constant acceleration for an infinite amount of time guarantees the existence of a temperature [17]. A well known example illustrating this fact is uniform circular motion [18]. We have shown that the Minkowski vacuum is perceived as a pure single-mode squeezed state by these NUAOs. In this case there is no need for an equivalent tracing over inaccessible degrees of freedom, or even a 38 motivation for doing such a thing given that this observer sees no event horizon and has complete knowledge of all information over the entire spacetime. Therefore, defining a temperature T = ~/2πkα is not justified despite the fact that the Bose-Einstein nature of the particle number distribution appears to suggest a thermal state. Thermal states should not be associated with accelerations in general, but only with accelerations capable of introducing horizons in the frame of the accelerating observer. This seems to suggest that the information paradox in black hole physics is an artifact of interpreting measurements of radiation seen by an unphysical stationary observer just outside the event horizon of a black hole. The observers considered in this thesis are in a sense “intermediate” between freely falling and stationary observers. These observers should shed light on the issues of black hole complementarity, firewalls [15, 16], and the information paradox. REFERENCES [1] W. Rindler, “Kruskal space and the Uniformly Accelerated Frame”, Am. J. Phys. 34, 1174-1178 (1966). [2] L. C. B. Crispino, A. Higuchi, and G. Matsas, “The Unruh effect and its applications”, Rev. Mod. Phys, 80 787-838 (2008). [3] E. Kasner, “Solutions of the Einstein Equations Involving Functions of Only One Variable”, Trans. Amer. Math. Soc., 27 155-162 (1925). [4] E. A. Milne, “World-Structure and the Expansion of the Universe”, Zeitschrift für Astrophysik, 27 1-95 (1933). [5] W. O. Kermack and W. H. McCrea, “On Milne’s Theory of World Structure”, Monthly Notices of the Royal Astronomical Society, 93 519-529 (1933). [6] S. W. Hawking, “Black-hole explosions”, Nature (London) 248, 3031 (1974); “Particle creation by black-holes”, Commun. Math. Phys. 43, 199220 (1975). [7] J. D. Bekenstein, “Black Holes and Entropy”, Phys. Rev. D 7, 2333 (1973); “Generalized second law of thermodynamics in black-hole physics”, Phys. Rev. D 9, 3292 (1974). [8] P. C. W. Davies, Particle production in Schwarzschild and Rindler metrics, J. Phys. A: Math Gen. 8, 609-616 (1975). [9] W. G. Unruh, “Notes on black-hole evaporation”, Phys. Rev. D 14, 870-892 (1976). [10] S. A. Fulling, “Nonuniqueness of canonical field quantization in Riemannian space-time”, Phys. Rev. D 7, 2850 (1973). [11] A. Doria and G. Muñoz, “Acceleration without Horizons”, arXiv:1502.05093 [gr-qc] (2015). [12] A. Doria and G. Muñoz, “Acceleration without Temperature”, arXiv:1503.01152 [gr-qc] (2015). [13] G. N. Watson, A Treatise on the Theory of Bessel Functions, 2nd ed. (Cambridge University Press, 1966). 40 [14] G. S. Agarwal, Quantum Optics (Cambridge University Press, 2013). [15] A. Almheiri, D. Marolf, J. Polchinski, and J. Sully, “Black Holes: Complementarity or Firewalls?”, J. High Energy Phys. 02, 062 (2013). [16] Y. Nomura, J. Varela, and S. Weinberg, “Black holes or firewalls: A theory of horizons”, Phys. Rev. D 88, 084052 (2013). [17] S. Deser and O. Levin, “Mapping Hawking into Unruh thermal properties”, Phys. Rev. D 59, 064004 (1999). [18] J. R. Letaw and J. D. Pfautseh, “Quantum scalar field in rotating coordinates”, Phys. Rev. D 22, 1345 (1980). 41 FIGURE t X =X1 X = X2 4 T =T 2 2 T =T 1 -4 2 -2 4 -2 -4 Figure 1. The spacetime diagram for a non-uniformly accelerating observer with coordinates (T, X) plotted with the corresponding Minkowski coordinates (t, x) light cone. The hyperbolae cross the light cone making explicit the absence of horizons. x APPENDICES Appendix A: Rindler Observers and Unruh Temperature We have already noted that equation (93) cannot be interpreted as the temperature of the particle distribution detected by our accelerating observer. Of course, a similar definition is warranted in the case of a Rindler observer. This naturally raises the question as to whether the Unruh temperature can be obtained from our results. At first sight it may appear that we have two incompatible expressions, since the Rindler observer is the α → 0 limit of our NUAO. However, as pointed out in [11], the limit should be taken as the NUAO’s incoming velocity v∞ → 1 with the product αγ∞ fixed, where γ∞ is the gamma factor associated with v∞ . Furthermore, in this limit the relationship between the observer’s time dT and the proper time dτ becomes γ∞ dT = dτ . This implies that energies and temperatures defined with respect to the proper time must include a gamma factor; in particular, Tτ = ~ 2πkαγ∞ (105) replaces equation (93). Since αγ∞ = 1/apr , as v∞ → 1 [11], the temperature becomes Tτ = ~ apr 2πk and we have agreement with the Unruh temperature. (106) 43 Appendix B: Verification of Bose-Einstein Distribution of Particle Numbers We check our expansion of the vacuum state by an independent calculation of the expectation value for the number operator. From equation (102), h0M |NΩj |0M i is h0M |Z −1/2 X Y (2ki − 1)!! −ki πΩi α p e NΩj |{2ki }i (2ki )! i {ki } (107) The number operator acting on the states gives NΩj |{2ki }i = 2kj |{2ki }i. Substitution of h0M |{2ki }i from (97) then yields h0M |NΩj |0M i = Z −1 2 X Y (2ki − 1)!! −ki πΩi α p e 2kj (2ki )! i {ki } (108) Equations (100) and (101) imply that the infinite product will produce a cancellation of every Zi factor in the product, with a corresponding term in the normalization, except when i = j. Hence h0M |NΩj |0M i = Zj−1 ∞ X [(2kj − 1)!!]2 −2kj πΩi α e 2kj (2k j )! k =0 (109) j The remaining sum is easily evaluated by taking a derivative of (100). The result is h0M |NΩj |0M i = (e2παΩj − 1)−1 This is the same Bose-Einstein distribution found in Section 2. (110) Fresno State Non-Exclusive Distribution License (to archive your thesis/dissertation electronically via the library’s eCollections database) By submitting this license, you (the author or copyright holder) grant to Fresno State Digital Scholar the non-exclusive right to reproduce, translate (as defined in the next paragraph), and/or distribute your submission (including the abstract) worldwide in print and electronic format and in any medium, including but not limited to audio or video. You agree that Fresno State may, without changing the content, translate the submission to any medium or format for the purpose of preservation. You also agree that the submission is your original work, and that you have the right to grant the rights contained in this license. You also represent that your submission does not, to the best of your knowledge, infringe upon anyone’s copyright. If the submission reproduces material for which you do not hold copyright and that would not be considered fair use outside the copyright law, you represent that you have obtained the unrestricted permission of the copyright owner to grant Fresno State the rights required by this license, and that such third-party material is clearly identified and acknowledged within the text or content of the submission. If the submission is based upon work that has been sponsored or supported by an agency or organization other than Fresno State, you represent that you have fulfilled any right of review or other obligations required by such contract or agreement. Fresno State will clearly identify your name as the author or owner of the submission and will not make any alteration, other than as allowed by this license, to your submission. By typing your name and date in the fields below, you indicate your agreement to the terms of this distribution license. Embargo options (fill box with an X). X Make my thesis or dissertation available to eCollections immediately upon submission. Embargo my thesis or dissertation for a period of 2 years from date of graduation. Embargo my thesis or dissertation for a period of 5 years from date of graduation. Alaric Doria Type full name as it appears on submission April 22, 2016 Date