TEMPERED FRACTIONAL BROWNIAN MOTION By Farzad Sabzikar A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Statistics – Doctor of Philosophy 2014 ABSTRACT TEMPERED FRACTIONAL BROWNIAN MOTION By Farzad Sabzikar Tempered fractional Brownian motion TFBM modifies the power law kernel in the moving average representation of a fractional Brownian motion (FBM), adding an exponential tempering. It also has a harmonizable representation. The increments of TFBM are stationary, and the autocovariance of the resulting tempered fractional Gaussian noise TFGN has semi-long range dependence, in which the autocorrelations decay like a power law over a moderate length scale, but eventually fall off more rapidly. TFBM can be represented as the linear combination of tempered fractional derivative (or tempered fractional integral) of the indicator functions. This representation and the classical Itˆo isometry provides to characterize the class of all deterministic functions for which the stochastic integral with respect to TFBM is well defined. Replacing the Gaussian random measure (Brownian motion) in the moving average or harmonizable representation of TFBM by a stable random measure, a linear tempered fractional stable motion (LTFSM), or a real harmonizable tempered fractional stable motion (HTFSM), respectively. Unlike the Gaussian case, LTFSM and HTFSM are two completely different processes. Existence, basic properties, sample path behavior, and dependence structure of both processes will be described. Keywords: Fractional Brownian motion, tempered fractional derivative, harmonizable representation, long range dependence, reproducing kernel Hilbert space. I dedicate this dissertation to my lovely wife, Sara Hazinia. iv ACKNOWLEDGMENTS I wish to express my deepest gratitude to my advisor Dr. Mark M. Meerschaert. His guidance, patience and thoughtfulness made this work possible. It was a privilege to work with him. I would like to thank my thesis committee members Drs. Hira Koul, V. S. Mandrekar, Yimin Xiao and Mantha S. Phanikumar, for their time and interest. I appreciate the help and support of the Department of Statistics and Probability in last five years. Finally, I thank my wife, Sara Hazinia for being there for me during this journey and my parents for their unconditional support. v TABLE OF CONTENTS LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 Chapter 2 Tempered Fractional Brownian Motion 2.1 Moving average representation . . . . . . . . . . 2.2 Harmonizable representation . . . . . . . . . . . . 2.3 Tempered fractional Gaussian noise . . . . . . . . 2.4 Sample path properties . . . . . . . . . . . . . . 2.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 4 8 9 13 15 Chapter 3 Tempered Fractional Calculus . . . 3.1 Tempered fractional calculus . . . . . . . . . 3.2 Stochastic Integrals . . . . . . . . . . . . . . 3.2.1 Semi-long range dependence . . . . . 3.2.2 Anti-persistence . . . . . . . . . . . . 3.2.3 Harmonizable representation . . . . . 3.3 Discussion . . . . . . . . . . . . . . . . . . . 3.3.1 White noise approach . . . . . . . . . 3.3.2 Reproducing kernel Hilbert space . . 3.3.3 Tempered distributions as integrands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 17 29 35 40 46 49 50 50 52 Chapter 4 Tempered fractional stable motion . . 4.1 Moving average representation . . . . . . . . . . 4.2 Tempered fractional harmonizable stable motion 4.3 Sample path properties . . . . . . . . . . . . . . 4.4 Local Times and Local nondeterminism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 54 72 76 81 Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 vi . . . . . . . . . . LIST OF FIGURES Figure 2.1 Figure 2.2 Figure 2.3 Figure 4.1 The autocovariance function (2.16) for TFGN with σ = 1, λ = 0.001 and H = 0.7 (solid line) and for the corresponding FGN with σ = 1, λ = 0 and H = 0.7 (dotted line). . . . . . . . . . . . . . . . . . . . 11 Left panel: Sample paths of TFBM (thick black line) with λ = 0.03 and H = 0.3, and FBM (thin black line) with H = 0.3. Both graphs use the same noise realization B(t). The right panel shows the same plots for λ = 0.01 and H = 0.7. . . . . . . . . . . . . . . . . . . . . . 15 The spectral density (2.21) for TFGN with σ = 1, λ = 0.06 and H = 0.7 (solid line) and FGN with σ = 1, λ = 0 and H = 0.7 (dotted line). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 Left panel: Sample paths of LTFSM (thick black line) with λ = 0.03 and H = 0.3, and LFSM (thin black line) with H = 0.3. Both graphs use the same noise realization Zα (t). The right panel shows the same plots for λ = 0.001, H = 0.7 and α = 1.5. . . . . . . . . . . . . . . . 77 vii Chapter 1 Introduction Fractional Brownian motion (FBM) is a Gaussian stochastic process whose increments, termed fractional Gaussian noise (FGN), can exhibit long range dependence [7, 22]. FBM has become popular in applications to science and engineering, since it yields a simple tractable model that captures the correlation structure seen in many natural systems [34, 43]. Formed by convolving Brownian motion with a power law, FBM is essentially the fractional integral or derivative of that Brownian motion [48, 56]. This is also reflected in the correlation function of FGN, which falls off like a power law with lag, and the corresponding spectral density, which behaves like a power law near the origin. The increments of long range dependent FBM have a spectral density that blows up like a power law at the origin. This diverging spectral density is one of the hallmarks of long range dependence. For wind speed measurements, the spectral density follows this power law model for moderate frequencies, but the data deviates from that model at low frequencies, and the measured spectal density remains bounded [55, 28, 53]. In Chapter 2 we define a closely related process, which we call tempered fractional Brownian motion (TFBM). It is defined by exponentially tempering the power law kernel in the moving average representation of a fractional Brownian motion. TFBM is a Gaussian process with stationary increments, and we call those increments tempered fractional Gaussian noise (TFGN). When FGN is long range dependent, TFGN exhibits semi-long range dependence. Its autocovariance function closely resembles that of FGN on an intermediate scale, but then 1 it eventually falls off more rapidly. Its spectral density resembles a negative power law for low frequencies, but eventually converges to zero at very low frequencies. TFBM can be a useful stochastic model for applications where the data follows FBM at some intermediate scale, but then deviates from FBM at longer scales. For example, wind speed measurements typically resemble long range dependent FBM over a range of frequencies, but deviate significantly at very low frequencies (corresponding to very long spatial scales). Since the spectal density of semi-long range dependent TFGN follows the same pattern, it can provide a useful model for such data. In Chapter 3 we develop the theory of stochastic integration for TFBM. Our approach follows the seminal work of Pipiras and Taqqu [56] for FBM. A FBM is the fractional derivative (or integral) of a Brownian motion, in a sense made precise in [56]. A fractional derivative (or integral) is a (distributional) convolution with a power law [48, 54, 60]. Multiplying that power law by an exponential factor leads to tempered fractional derivatives and integrals. TFBM can be written in terms of tempered fractional derivatives (or integrals) of a Brownian motion. This representation and the classical Itˆo Isometry allows us to characterize the class of all deterministic functions for which the stochastic integral with respect to TFBM is well defined. We also apply the harmonizable representation of TFBM to define another type of stochastic integral of deterministic functions with respect to TFBM. In chapter 4 we consider heavy tailed analogues to TFBM. Replacing the Gaussian random measure (Brownian motion) in the moving average or harmonizable representation of TFBM by a stable random measure, we obtain a linear tempered fractional stable motion (LTFSM), or a real harmonizable tempered fractional stable motion (HTFSM), respectively. Unlike the Gaussian case, LTFSM and HTFSM are two completely different processes. Existence, basic properties, sample path behavior, and dependence structure of both processes 2 are described in this thesis. We also prove that LTFSM and HTFSM are locally nondeterministic on every compact interval. 3 Chapter 2 Tempered Fractional Brownian Motion This Chapter has five sections. In Section 2.1, we define tempered fractional Brownian motion (TFBM) using a moving average representation, and we establish some of its basic properties. Section 2.2 develops the harmonizable representation of TFBM, and Section 2.3 discusses tempered fractional Gaussian noise (TFGN). Sample path properties of TFBM are proven in Section 2.4, and an application to wind speed is discussed in Section 2.5. 2.1 Moving average representation Let {B(t)}t∈R be a real-valued Brownian motion on the real line, a process with stationary independent increments such that B(t) has a Gaussian distribution with mean zero and variance σ 2 |t| for all t ∈ R, for some σ > 0. Define an independently scattered Gaussian random measure B(dx) with control measure m(dx) = σ 2 dx by setting B[a, b] = B(b)−B(a) for any real numbers a < b, and then extending to all Borel sets. Then the stochastic integrals ∫ ∫ I(f ) := f (x)B(dx) are defined for all functions f : R → R such that f (x)2 dx < ∞, as ∫ Gaussian random variables with mean zero and covariance E[I(f )I(g)] = σ 2 f (x)g(x)dx, see for example Chapter 3 in [61]. Definition 2.1.1. Given an independently scattered Gaussian random measure B(dx) on R 4 with control measure σ 2 dx, for any α < 12 and λ ≥ 0, the stochastic integral ∫ +∞ [ ] −λ(−x)+ (−x)−α B(dx), Bα,λ (t) := e−λ(t−x)+ (t − x)−α − e + + −∞ (2.1) where (x)+ = xI(x > 0), and 00 = 0, will be called a tempered fractional Brownian motion (TFBM). It is easy to check that the function −λ(−x)+ (−x)−α gα,λ,t (x) := e−λ(t−x)+ (t − x)−α + −e + (2.2) is square integrable over the entire real line for any α < 21 , so that TFBM is well-defined, and of course FBM is a special case of TFBM with λ = 0. Note also that gα,λ,ct (cx) = c−α gα,cλ,t (x), (2.3) for all t, x ∈ R and all c > 0. The next result shows that TFBM has a nice scaling property, involving both the time scale and the tempering. Here the symbol indicates equality of finite dimensional distributions. Proposition 2.1.2. TFBM (2.1) is Gaussian stochastic process with stationary increments, such that { } Bα,λ (ct) t∈R { } H c Bα,cλ (t) t∈R (2.4) for any scale factor c > 0, where the Hurst index H = 1/2 − α. Proof. Since B(dx) has control measure m(dx) = σ 2 dx, the random measure B(c dx) has 5 control measure c1/2 σ 2 dx. Given t1 < t2 < · · · < tn , a change of variable x = cx′ then yields ( ) Bα,λ (cti ) : i = 1, . . . , n = (∫ (∫ ) gα,λ,cti (x)B(dx) : i = 1, . . . , n c−α gα,cλ,ti (x′ )c1/2 B(dx′ ) : i = 1, . . . , n ) so that (2.4) holds with H = 1/2 − α. For any s, t ∈ R, the integrand (2.2) satisfies gα,λ,s+t (s + x) − gα,λ,s (s + x) = gα,λ,t (x), and hence a change of variable x = s + x′ in the moving average representation yields ( Bα,λ (s + ti ) − Bα,λ (s) : i = 1, . . . , n (∫ ) gα,λ,ti (x′ )B(dx′ ) : i = 1, . . . , n ) which shows that TFBM has stationary increments. Proposition 2.1.3. The covariance function of TFBM (2.1) has the form ] [ ] σ 2 [ 2 2H 2 |t − s|2H , Ct |t| + Cs2 |s|2H − Ct−s Cov Bα,λ (t), Bα,λ (s) = 2 (2.5) for any s, t ∈ R, where H = 1/2 − α. Here Ct2 = 2Γ(H + 12 ) 2Γ(2H) 1 √ − KH (λ|t|), 2H π (2λ|t|) (2λ|t|)H (2.6) for t ̸= 0, Ct2 = 0 when t = 0, where Kν (z) is the modified Bessel function of the second kind. 6 Proof. Use the moving average representation (2.1) with σ = 1 to define Ct2 : = E[Bα,λ|t| (1)2 ] = ∫ +∞ = −∞ −2 ∫ +∞ [ ] −λt(−x)+ (−x)−α 2 dx e−λt(1−x)+ (1 − x)−α − e + + −∞ e−2λt(1−x)+ (1 − x)−2α + dx + ∫ +∞ −∞ ∫ +∞ −∞ e−2λt(−x)+ (−x)−2α + dx (2.7) −λt(−x)+ (−x)−α dx. e−λt(1−x)+ (1 − x)−α + e + Apply the definition of the gamma function, along with a standard integral formula from Page 344 in [24], to see that (2.6) holds. Since TFBM has stationary increments, it follows from (2.4) that E[Bα,λ (t)2 ] = |t|2H Ct2 for all t real. Recall the elementary formula ab = 1 2 2 2 α,λ (t) and b = B α,λ (s), take expectations, and use the 2 [a + b − (a − b) ], set a = B stationary increments property again, to see that (2.5) holds. Remark 2.1.4. The integral representation (2.1) is causal, i.e., Bα,λ (t) depends only on the values of B(s) for s ≤ t. For applications to spatial statistics, consider p,q Bα,λ (t) = p ∫ +∞ [ −∞ ∫ +∞ [ +q −∞ −λ(−x)+ (−x)−α e−λ(t−x)+ (t − x)−α + −e + −λ(x)+ (x)−α e−λ(x−t)+ (x − t)−α + −e + ] B(dx) (2.8) ] B(dx) for p, q ≥ 0. It is not hard to check, by mimicking the proof of Proposition 2.1.2, that this process also has stationary increments, and satisfies the scaling property { } p,q Bα,λ (ct) t∈R { } p,q cH Bα,cλ (t) t∈R (2.9) for any scale factor c > 0, where the Hurst index H = 1/2 − α. When p = q > 1, (2.8) is a well-balanced TFBM. 7 2.2 Harmonizable representation ˆ1 and B ˆ2 be independent Gaussian random measures with B ˆ1 (A) = B ˆ1 (−A), B ˆ2 (A) = Let B ˆ2 (−A) and E[(B ˆi (A))2 ] = m(A)/2, where m(dx) = σ 2 dx, and define the complex-valued −B ˆ=B ˆ1 + iB ˆ2 . If f (x) is a complex-valued function of x real such Gaussian random measure B ∫ ∫ that its Fourier transform fˆ(k) := (2π)−1/2 e−ikx f (x) dx exists and |fˆ(k)|2 dk < ∞, ∫ ∫ ∫ ˆ fˆ) = fˆ(k)B(dk) ˆ ˆ1 (dk) − fˆ2 (k)B ˆ2 (dk), we define the stochastic integral I( := fˆ1 (k)B ˆ fˆ) is a Gauswhere fˆ = fˆ1 + ifˆ2 is separated into real and imaginary parts. Then I( ∫ ˆ fˆ)I(ˆ ˆ g )] = fˆ(k)ˆ sian random variable with mean zero, such that E[I( g (k) dk. The Par∫ ∫ ∫ ∫ seval identity f (x)g(x) dx = fˆ(k)ˆ g (k) dk implies that ( f (x)B(dx), g(x)B(dx)) ∫ ∫ ˆ ˆ ( fˆ(k)B(dk), gˆ(k)B(dk)), see Proposition 7.2.7 in [61]. Proposition 2.2.1. The TFBM (2.1) has the harmonizable representation Γ(1 − α) B α,λ (t) = √ 2π ∫ +∞ −itk e −1 ˆ B(dk). 1−α −∞ (λ − ik) (2.10) Proof. To show that the stochastic integral (2.10) exists, note that ∫ +∞ −∞ 2 ∫ +∞ e−itx − 1 4 dx < ∞, dx ≤ 1−α 2 2 1−α (λ − ix) −∞ (λ + x ) since the last integrand is bounded and O(x2α−2 ) as |x| → ∞, for α < 21 . Observe that the 8 function gα,λ,t , given by (2.2), has the Fourier transform ∫ +∞ [ ] 1 −λ(−x)+ (−x)−α dx gα,λ,t (k) = √ e−ikx e−λ(t−x)+ (t − x)−α − e + + 2π −∞ [∫ t ] ∫ 0 1 −ikx −λ(t−x) −α −ikx λx −α =√ e e (t − x) dx − e e (−x) dx 2π −∞ −∞ [ ] ∫ +∞ ∫ +∞ 1 −ikt −u(λ−ik) −α −u(λ−ik) −α =√ e e u du − e u du 2π 0 0 Γ(1 − α) e−ikt − 1 . = √ 2π (λ − ik)1−α (2.11) Hence by (2.1), ∫ +∞ Bα,λ (t) = −∞ ∫ +∞ gα,λ,t (x)B(dx) ∫ Γ(1 − α) +∞ e−ikt − 1 ˆ ˆ gα,λ,t (k)B(dk) = √ B(dk), 1−α 2π −∞ −∞ (λ − ik) which is equivalent to (2.10). Remark 2.2.2. The spectral representation (2.10) reduces to that of causal FBM in the special case λ = 0, see for example Equation 7.2.17 in [61]. The general TFBM (2.8) has spectral representation [ ] ∫ −itk Γ(1 − α) e −1 p ik q ik p,q Bα,λ (t) = √ − B(dk). ik (λ − ik)1−α (λ + ik)1−α 2π R 2.3 (2.12) Tempered fractional Gaussian noise Given a TFBM (2.1), we define tempered fractional Gaussian noise (TFGN) Xj = Bα,λ (j + 1) − Bα,λ (j), 9 for integers −∞ < j < ∞. (2.13) It follows easily from (2.1) that TFGN has the moving average representation ∫ +∞ [ ] −λ(j−x)+ (j − x)−α B(dx). Xj = e−λ(j+1−x)+ (j + 1 − x)−α − e + + −∞ (2.14) Using (2.10), it also follows that the harmonizable representation of TFGN is ∫ Γ(1 − α) +∞ −ikj e−ik − 1 ˆ Xj = √ e B(dk). (λ − ik)1−α 2π −∞ (2.15) It follows from (2.5) that TFGN is a stationary Gaussian time series with mean zero and covariance function r(j) := E[X0 Xj ] = ] σ2 [ 2 − 2 |j|2H C 2 + |j − 1|2H C 2 |j + 1|2H Cj+1 j j−1 , 2 (2.16) where H = 1/2 − α, and Cj is given by (2.6). Remark 2.3.1. Using the well-known fact that Kν (x) ∼ √ π(2x)−1/2 e−x as x → ∞, it follows easily from (2.6) that t2H Ct2 → 2Γ(2H)(2λ)−2H , as t → ∞. (2.17) Hence Cj ∼ Cj+1 as j → ∞. Then, (2.16) along with a Taylor series expansion, shows that r(j) ∼ σ 2 Cj2 H(2H − 1)|j|2H−2 as j → ∞. Compare this with Proposition 7.2.10 in [61]. For λ > 0, sufficiently small, the power law terms in (2.7) dominate, Cj2 remains almost constant, and r(j) falls off like |j|2H−2 10 for moderate values of j > 0. For larger j, the exponential terms in (2.7) dominate, and (2.17) implies that r(j) ∼ j −2 2H(2H − 1)Γ(2H)(2λ)−2H , as j → ∞. Hence TFGN is short range dependent, but its covariance function is arbitrarily close to that of long range dependent FGN for small values of λ, and moderate lags. We call this property semi-long range dependence, since it is analogous to the semi-heavy tails of Barndorff and Nielsen [21]. Figure 2.1 shows a log-log plot of r(j) in the case H = 0.7 and λ = 0.001, where FGN exhibits long range dependence. 1 0.1 0.01 0.001 1 10 100 lag Figure 2.1: The autocovariance function (2.16) for TFGN with σ = 1, λ = 0.001 and H = 0.7 (solid line) and for the corresponding FGN with σ = 1, λ = 0 and H = 0.7 (dotted line). Proposition 2.3.2. TFGN (2.13) has the spectral density h(k) = +∞ 2 ∑ Γ(1 − α)2 −ik σ2 . e −1 2π 2 + (k + 2πℓ)2 ]H+1/2 [λ ℓ=−∞ (2.18) Proof. Recall that the spectral density ∫ π +∞ 1 ∑ ikj h(k) = e r(j) and r(j) = e−ikj h(k)dk. 2π −π j=−∞ 11 (2.19) Define C = √ 2π/Γ(1 − α) and apply (2.15) to write 2 −ik − 1 ∫ σ 2 +∞ −ikj e r(j) = 2 e dk C −∞ (λ2 + k 2 )(1−α) ∫ +π +∞ 2 ∑ 1 σ2 −ikj −ik = 2 e e −1 dk (1−α) C −π [λ2 + (k + 2πℓ)2 ] (2.20) ℓ=−∞ and then it follows from (2.19) that the spectral density of TFGN is given by (2.18). Remark 2.3.3. Extending (2.13) to all j real, we obtain the continuous parameter TFGN Xt = Bα,λ (t + 1) − Bα,λ (t). The harmonizable representation of this process is given by (2.15) with j replaced by t, and the proof of Proposition 2.3.2 implies that Xt has spectral density h(ω) = 2 Γ(1 − α)2 −iω σ2 e −1 H+1/2 2π [λ2 + ω 2 ] (2.21) for all real ω. The fact that e−iω −1 ∼ −iω as ω → 0 yields the low frequency approximation σ 2 Γ(1 − α)2 ω2 h(ω) ≈ . H+1/2 2π (λ2 + ω 2 ) See Section 2.5 for an application to wind speed data. 12 2.4 Sample path properties We say that the sample paths of a stochastic process X(t) satisfy a uniform H¨older condition of order β on the compact set K ⊂ R if there exists a positive random variable A such that |X(x) − X(y)| ≤ A|x − y|β , almost surely for all x, y ∈ K. We say that the process has H¨older critical exponent γ ∈ (0, 1) if the process satisfies a uniform H¨older condition of any order β ∈ (0, γ) on any compact set K ⊂ R, and fails to satisfy this condition for β ∈ (γ, 1). Theorem 2.4.1. The sample paths of the TFBM (2.1) have H¨older critical exponent H = 1/2 − α, for any α ∈ (−1/2, 1/2), and for any λ ≥ 0. Proof. Since Bα,λ (0) = 0, it follows from Proposition 4 in [10] that if { [ ] ( ) γ = sup β > 0 : E Bα,λ (t)2 = o |t|2β } as |t| → 0 , (2.22) then the TFBM Bα,λ (t) satisfies a uniform H¨older condition of order β on any compact set for any β ∈ (0, γ). Moreover, if we also have { ( [ ]) γ = inf β > 0 : |t|2β = o E Bα,λ (t)2 } as |t| → 0 , (2.23) then this TFBM has H¨older critical exponent γ. Use the harmonizable representation (2.10) 13 to write [ −itk ] ∫ +∞ −itk [ ] 1 e −1 e −1 2 dk E Bα,λ (t) = 2 C −∞ (λ − ik)1−α (λ − ik)1−α ∫ +∞ 2 = 2 [1 − cos(tk)] (λ2 + k 2 )α−1 dk, C −∞ √ 2π/Γ(1−α), and apply the Tauberian theorem for Fourier transforms, Theorem [ ] 1 in [57], to see that E Bα,λ (t)2 ∼ H(1/t) as t → 0, where where C = ∫ 2 H(x) = 2 (λ2 + k 2 )α−1 dk. C |k|>x Since λ2 +k 2 ∼ k 2 as k → ∞, for any ε > 0, there exists some M > 0 such that (1−ε)k 2α−2 < (λ2 + k 2 )α−1 < (1 + ε)k 2α−2 for all k > M , and hence we have 4(1 − ε) 2α−1 < H(x) < 4(1 + ε) x2α−1 x (1 − 2α)C 2 (1 − 2α)C 2 , for all x > M . Substitute t = 1/x to see that both (2.22) and (2.23) hold with γ = 1−2α = 2H, which completes the proof. Remark 2.4.2. The harmonizable representation ∫ +∞ ( X(t) = −∞ ) ˆ e−itk − 1 fˆ(k)B(dk) defines a mean zero Gaussian processes with stationary increments for any Fourier filter fˆ(k) ∫ such that [1 − cos(tk)] |fˆ(k)|2 dk < ∞. If |fˆ(k)|2 is regularly varying at infinity with index 2α − 2 for some −1/2 < α < 1/2, the Karamata Theorem (e.g., see Lemma 5.3.8 (d) in [47]) implies that H(x) varies regularly at infinity with index 2α − 1, and then the proof 14 X(t) 0 10 20 10 5 −5 −20 0 X(t) 0 100 200 300 400 500 0 t 100 200 300 400 500 t Figure 2.2: Left panel: Sample paths of TFBM (thick black line) with λ = 0.03 and H = 0.3, and FBM (thin black line) with H = 0.3. Both graphs use the same noise realization B(t). The right panel shows the same plots for λ = 0.01 and H = 0.7. of Theorem 2.4.1 extends to show that X(t) has H¨older critical exponent 1 − 2α. Several examples of such processes are given in [10]. The sample paths of TFBM closely resemble that of FBM for small values of the tempering parameter λ > 0. The left panel in Figure 2.2 compares a typical sample path of both processes, simulated using the same white noise B(dx), in a case where FBM is negative dependent. The right panel shows the corresponding sample paths in a case where FBM is long range dependent. These simulations use a discretized version of the moving average representation (2.1). It would also be interesting to develop a simulation method based on the harmonizable representation (2.10). 2.5 Discussion Wind speed data are important for electrical power generation and structural engineering. The most popular model for wind speed near the earth surface, due to Davenport [16], see also [38], can be written in the form st = µ + Xt , where µ = E[st ] is the average wind speed, 15 0.3 0.25 0.2 0.15 0.1 0.05 0 0 0.5 1 1.5 2 frequency Figure 2.3: The spectral density (2.21) for TFGN with σ = 1, λ = 0.06 and H = 0.7 (solid line) and FGN with σ = 1, λ = 0 and H = 0.7 (dotted line). and Xt has normalized spectral density 4800DV10 x2 4 (1 + x2 ) 3 , (2.24) where V10 is the mean velocity (m/sec) at an altitude of 10 meters, D is the corresponding drag coefficient, and x = 1200ω/V10 . In view of Remark 2.3.3, it is not hard to check that (2.24) corresponds to the the spectral density of a continuous parameter TFGN with λ = V10 /1200 and H = 5/6. Hence TFGN can provide a useful stochastic process model for wind speed data. Figure 2.3 compares the spectral density of TFGN and FGN in the case where FGN is long range dependent. The spectral density of FGN blows up at the origin like a power law. The spectral density of TFGN follows the same power law at moderate frequencies, but remains bounded at very low frequencies, a behavior typically seen in wind speed data for example in [55, 16, 53, 28]. 16 Chapter 3 Tempered Fractional Calculus This chapter has three sections. In Section 3.1 we prove some basic results on tempered fractional calculus, which will be needed in the sequel. In Section 3.2 we apply the methods of Section 3.1 to construct a suitable theory of stochastic integration for tempered fractional Brownian motion. Finally, in Section 3.3 we discuss model extensions, related results, and some open questions. 3.1 Tempered fractional calculus In this section, we define tempered fractional integrals and derivatives, and establish their essential properties. These results will form the foundation of the stochastic integration theory developed in Section 3.2. We begin with the definition of a tempered fractional integral. Definition 3.1.1. For any f ∈ Lp (R) (where 1 ≤ p < ∞), the positive and negative tempered fractional integrals are defined by and ∫ +∞ 1 α,λ −λ(t−u)+ du, I+ f (t) = f (u)(t − u)α−1 + e Γ(α) −∞ (3.1) ∫ ∞ 1 α,λ I− f (t) = e−λ(u−t)+ du, f (u)(u − t)α−1 + Γ(α) −∞ (3.2) 17 ∫ respectively, for any α > 0 and λ > 0, where Γ(α) = 0+∞ e−x xα−1 dx is the Euler gamma function, and (x)+ = xI(x > 0). When λ = 0 these definitions reduce to the (positive and negative) Riemann-Liouville fractional integral [48, 54, 60], which extends the usual operation of iterated integration to a fractional order. When λ = 1, the operator (3.1) is called the Bessel fractional integral [60, Section 18.4]. α,λ Lemma 3.1.2. For any α > 0, λ > 0, and p ≥ 1, I± is a bounded linear operator on Lp (R) such that ∥I± f ∥p ≤ λ−α ∥f ∥p , α,λ (3.3) for all f ∈ Lp (R). Proof. Young’s Theorem [60, p. 12] states that if ϕ ∈ L1 (R) and f ∈ Lp (R) then ϕ∗f ∈ Lp (R) and the inequality ∥ϕ ∗ f ∥p ≤ ∥ϕ∥1 ∥f ∥p , (3.4) where ∗ denotes the convolution [f ∗ ϕ](t) = ∫ +∞ −∞ f (u)ϕ(t − u)du = [ϕ ∗ f ](t). Obviously I± is linear, and I± f (t) = [f ∗ ϕ± α ](t) where α,λ α,λ ϕ+ α (t) = 1 α−1 −λt t e 1(0,∞) (t), Γ(α) 1 ϕ− (−t)α−1 e−λ(−t) 1(−∞,0) (t) α (t) = Γ(α) 18 (3.5) for any α, λ > 0. But ∥ϕ± α ∥1 = ∫ +∞ ] 1 1 [ −α λ Γ(α) = λ−α , tα−1 e−λt dt = Γ(α) 0 Γ(α) using the formula for the Laplace transform (moment generating function) of the gamma probability density, and then (3.3) follows from Young’s Inequality (3.4). Next we prove a semigroup property for tempered fractional integrals, which follows easily from the following property of the convolution kernels in the definition (3.1.1). Lemma 3.1.3. For any λ > 0 the functions (3.5) satisfy ± ± ϕ± α ∗ ϕβ = ϕα+β , (3.6) for any α > 0 and β > 0. Proof. For t > 0 we have ∫ t 1 + + (t − s)α−1 e−λ(t−s) sβ−1 e−λs ds ϕα ∗ ϕβ (t) = Γ(α)Γ(β) 0 = 1 tα+β−1 e−λt = ϕ+ α+β (t) Γ(α + β) using the formula for the beta probability density. The proof for ϕ− α is similar. The following lemma establishes the semigroup property for tempered fractional integrals on Lp (R). In the case λ = 0, the semigroup property for fractional integrals is well known (e.g., see Samko et al. [60, Theorem 2.5]). 19 Lemma 3.1.4. For any λ > 0 we have α,λ β,λ α+β,λ I± I± f = I± f (3.7) for all α, β > 0 and all f ∈ Lp (R). Proof. Lemma 3.1.2 shows that both sides of (3.7) belong to Lp (R) for any f ∈ Lp (R), and α,λ then the result follows immediately from Lemma 3.1.3 along with the fact that I± f (t) = [f ∗ ϕ± α ](t). The next result shows that positive and negative tempered fractional integrals are adjoint ∫ operators with respect to the inner product ⟨f, g⟩2 = f (x)g(x) dx on L2 (R). Lemma 3.1.5 (Integration by parts). Suppose f, g ∈ L2 (R). Then α,λ α,λ ⟨f, I+ g⟩2 = ⟨I− f, g⟩2 (3.8) for any α > 0 and any λ > 0. Proof. Write ∫ +∞ −∞ ∫ x 1 g(u)(x − u)α−1 e−λ(x−u) du dx f (x) Γ(α) −∞ −∞ ∫ ∫ +∞ g(u) +∞ = f (x)(x − u)α−1 e−λ(x−u) dx du Γ(α) u −∞ ∫ +∞ α,λ = I− f (x)g(x) dx α,λ f (x)I+ g(x) dx = ∫ +∞ −∞ and this completes the proof. Next we discuss the relationship between tempered fractional integrals and Fourier trans20 forms. Recall that the Fourier transform ∫ +∞ 1 ˆ F[f ](k) = f (k) = √ e−ikx f (x)dx 2π −∞ , for functions f ∈ L1 (R) ∩ L2 (R) can be extended to an isometry (a linear onto map that preserves the inner product) on L2 (R) such that ∫ n 1 f (k) = lim √ e−ikx f (x) dx, n→∞ 2π −n (3.9) for any f ∈ L2 (R), see for example [31, Theorem 6.6.4]. Lemma 3.1.6. For any α > 0 and λ > 0 we have F[I± f ](k) = fˆ(k)(λ ± ik)−α , α,λ (3.10) for all f ∈ L1 (R) and all f ∈ L2 (R). Proof. The function ϕ+ α in (3.5) has Fourier transform F[ϕ+ α ](k) = ∫ ∞ 1 1 √ e−ikt tα−1 e−λt dt = √ (λ + ik)−α Γ(α) 2π 0 2π (3.11) by the formula for the Fourier transform of a gamma density. For any two functions f, g ∈ √ L1 (R), the convolution f ∗ g ∈ L1 (R) has Fourier transform 2π fˆ(k)ˆ g (k) (e.g., see [48, α,λ p. 65]), and then (3.10) follows. The argument for I− is quite similar. If f ∈ L2 (R), approximate by the L1 function f (x)1[−n,n] (x) and let n → ∞. Remark 3.1.7. Recall that the space of rapidly decreasing functions S(R) consists of the 21 infinitely differentiable functions g : R → R such that sup xn g (m) (x) < ∞, x∈R where n, m are non-negative integers, and g (m) is the derivative of order m. The space S ′ (R) of continuous linear functionals on S(R) is called the space of tempered distributions. The Fourier transform, and inverse Fourier transform, can then be extended to linear continuous mappings of S ′ (R) into itself. If f : R → R is a measurable function with polynomial growth, ∫ ∫ so that |f (x)|(1 + |x|)−p dx < ∞ for some p > 0, then Tf (φ) = f (x)φ(x) dx := ⟨f, φ⟩1 is a tempered distribution, also called a generalized function. The Fourier transform of this generalized function is defined as Tˆf (φ) = ⟨fˆ, φ⟩1 = ⟨f, φ⟩ ˆ 1 = Tf (φ) ˆ for φ ∈ S(R). See Yosida [68, Ch.VI] for more details. If f is a tempered distribution, then the tempered fractional integrals I± f (x) exist as convolutions with the tempered distributions ϕ± α in α,λ (3.5). The same holds for Riemann-Liouville fractional integrals (the case λ = 0), but that 1 case is more delicate, because the power law kernels ϕ± α of (3.5) with λ = 0 are not in L (R). Next we consider the inverse operator of the tempered fractional integral, which is called a tempered fractional derivative. For our purposes, we only require derivatives of order 0 < α < 1, and this simplifies the presentation. Definition 3.1.8. The positive and negative tempered fractional derivatives of a function f : R → R are defined as ∫ t α f (t) − f (u) −λ(t−u) α,λ α e du, D+ f (t) = λ f (t) + Γ(1 − α) −∞ (t − u)α+1 22 (3.12) and ∫ +∞ f (t) − f (u) −λ(u−t) e du, (u − t)α+1 α α,λ D− f (t) = λα f (t) + Γ(1 − α) t (3.13) respectively, for any 0 < α < 1 and any λ > 0. If λ = 0, the definitions (3.12) and (3.13) reduce to the positive and negative Marchaud fractional derivatives [60, Section 5.4]. Note that tempered fractional derivatives cannot be defined pointwise for all functions f ∈ Lp (R), since we need |f (t) − f (u)| → 0 fast enough to counter the singularity of the denominator (t − u)α+1 , as u → t. Next we establish the existence and compute the Fourier transform of tempered fractional derivatives on a natural domain. Theorem 3.1.9. Assume f and f ′ are in L1 (R). Then the tempered fractional derivative α,λ D+ f (t) exists and α,λ F[D± f ](k) = f (k)(λ ± ik)α , (3.14) for any 0 < α < 1 and any λ > 0. Proof. A standard argument from functional analysis (e.g., see [50, Proposition 2.2]) shows that if f, f ′ ∈ L1 (R), then ∫ ∫ I := |f (t) − f (u)| dt du < ∞, 1+α R R |t − u| 23 (3.15) for any 0 < α < 1. To see this, write I = I1 + I2 where ∫ ∫ |f (t) − f (u)| dt du 1+α R R∩{|t−u|<1} |t − u| ∫ ∫ |f (t) − f (z + t)| = dz dt |z|1+α R {|z|<1} ∫ ∫ ∫ 1 −α ≤ |z| |f ′ (t + uz)| du dz dt = I1 : = R {|z|<1} 0 2 ∥f ′ ∥L1 (R) < ∞ 1−α and ∫ ∫ |f (t) − f (u)| dt du 1+α R R∩{|t−u|≥1} |t − u| ∫ ∫ |f (t)| + |f (z + t)| 2 ≤ dt dz = ∥f ∥L1 (R) < ∞. 1+α α |z| R {|z|≥1} I2 : = Now it follows easily from (3.15) that D± f exists for all f, f ′ ∈ L1 (R). Define α,λ ∫ t f (t) − f (u) −λ(t−u) α e du, F (t) := Γ(1 − α) −∞ (t − u)α+1 and apply the Fubini Theorem, along with the shift property F[f (t − y)](k) = e−iky fˆ(k) of the Fourier transform, to see that ∫ +∞ ∫ ∞ α f (t) − f (t − y) −λy −ikt √ e e dy dt F (k) = y α+1 Γ(1 − α) 2π −∞ 0 ∫ +∞ ( ) α I (α) = f (k), y −α−1 e−λy 1 − e−iky f (k) dy = λ Γ(1 − α) 0 Γ(1 − α) where ∫ +∞ ( Iλ (α) = 0 ) e−λy − e−(λ+ik)y αy −α−1 dy. 24 (3.16) Integrate by parts with u = e−λy − e−(λ+ik)y and use the fact that e−λy − e−(λ+ik)y = O(y) as y → 0, to obtain [( Iλ (α) = e−λy − e−(λ+ik)y )( ) −y −α ∫ ∞ ] ∞ + 0 0 [ y −α −λe−λy + (λ + ik)e−(λ+ik)y ] dy. Use the definition of the gamma function, and the formula for the Fourier tranform of the gamma probability density, to obtain ∫ ∞ ∫ ∞ y −α e−(λ+ik)y dy 0 0 ( ) Γ(1 − α) ik α−1 α = −λ Γ(1 − α) + (λ + ik) 1−α 1+ λ λ Iλ (α) = −λ y −α e−λy dy + (λ + ik) = Γ(1 − α) [(λ + ik)α − λα ] . α,λ Then F (k) = f (k) [(λ + ik)α − λα ], and hence F[D+ f ](k) = (λ + ik)α fˆ(k). The proof for α,λ F[D− f ](k) is similar. Remark 3.1.10. Theorem 3.1.9 can also be proven, under somewhat stronger conditions, using the generator formula for infinitely divisible semigroups [48, Theorem 3.17 and Theorem 3.23 (b)]. Next we extend the definition of tempered fractional derivatives to a suitable class of functions in L2 (R). For any α > 0 and λ > 0, we define the fractional Sobolev space ∫ W α,2 (R) := {f ∈ L2 (R) : R (λ2 + k 2 )α |fˆ(k)|2 dk < ∞}, (3.17) which is a Banach space with norm ∥f ∥α,λ = ∥(λ2 + k 2 )α/2 fˆ(k)∥2 . The space W α,2 (R) is the same for any λ > 0 (typically we take λ = 1) and all the norms ∥f ∥α,λ are equivalent, 25 since 1 + k 2 ≤ λ2 + k 2 ≤ λ2 (1 + k 2 ) for all λ ≥ 1, and λ2 + k 2 ≤ 1 + k 2 ≤ λ−2 (1 + k 2 ) for all 0 < λ < 1. α,λ Definition 3.1.11. The positive (resp., negative) tempered fractional derivative D± f (t) of a function f ∈ W α,2 (R) is defined as the unique element of L2 (R) with Fourier transform f (k)(λ ± ik)α , for any α > 0 and any λ > 0. Remark 3.1.12. The pointwise definition of the tempered fractional derivative in real space is more complicated when α > 1. For example, when 1 < α < 2 we have ∫ t α f (u) − f (t) + (t − u)f ′ (t) −λ(t−u) α,λ α α−1 ′ f (x) + D+ f (t) = λ f (t) + αλ e du, Γ(1 − α) −∞ (t − u)α+1 for all f ∈ W 1,2 (R), compare [48, Remark 7.11]. Lemma 3.1.13. For any α > 0, β > 0 and λ > 0 we have α,λ β,λ α+β,λ D± D± f (t) = D± f (t), for any f ∈ W α+β,2 (R). Proof. It is obvious from (3.17) that W α,2 (R) ⊂ W β,2 (R) for α > β. It is clear from β,λ Definition 3.1.11 that D± f (t) exists and belongs to W α,2 (R) for any f ∈ W α+β,2 (R), and α,λ likewise, D± f (t) exists and belongs to L2 (R) for any f ∈ W α,2 (R). Lemma 3.1.14. For any α > 0 and λ > 0, we have α,λ α,λ D± I± f (t) = f (t) 26 (3.18) for any function f ∈ L2 (R), and α,λ α,λ I± D± f (t) = f (t) (3.19) for any f ∈ W α,2 (R). α,λ Proof. Given f ∈ L2 (R), note that g(t) = I± f (t) satisfies gˆ(k) = fˆ(k)(λ±ik)−α by Lemma 3.1.6, and then it follows easily that g ∈ W α,2 (R). Definition 3.1.11 implies that α,λ α,λ α,λ F[D± I± f ](k) = F[D± g](k) = g(k)(λ ± ik)α = fˆ(k), (3.20) and then (3.18) follows using the uniqueness of the Fourier transform. The proof of (3.19) is similar. Lemma 3.1.15. Suppose f, g ∈ W α,2 (R). Then α,λ α,λ ⟨f, D+ g⟩2 = ⟨D− f, g⟩2 , (3.21) for any α > 0 and any λ > 0. Proof. Apply the Plancherel Theorem along with Definition 3.1.11 to see that α,λ ⟨f, D+ g⟩2 = ∫ α,λ α,λ f (x)D+ g(x) dx = ⟨fˆ, (λ + ik)α gˆ⟩2 = ⟨(λ − ik)α fˆ, gˆ⟩2 = ⟨D− f, g⟩2 and this completes the proof. Remark 3.1.16. One can also prove (3.21) for f, f ′ , g, g ′ ∈ L1 (R) ∩ L2 (R) using integration by parts, compare [69, Appendix A.1]. 27 A slightly different tempered fractional derivative ∫ t α f (t) − f (u) −λ(t−u) α,λ D+ f (t) = e du, Γ(1 − α) −∞ (t − u)α+1 ∫ +∞ α f (t) − f (u) −λ(u−t) α,λ e du, D− f (t) = Γ(1 − α) t (u − t)α+1 (3.22) was proposed by Cartea and del-Castillo-Negrete [12] for a problem in physics, and studied further by Baeumer and Meerschaert [4, 48] using tools from probability theory and semiα,λ groups. When 0 < α < 1, F[D± f ](k) = f (k)[(λ ± ik)α − λα ]fˆ(k) for suitable functions f . The additional λα term makes the evolution equation ∂ α,λ α,λ u(x, t) = [pD+ + qD− ]u(x, t), ∂t (3.23) for p, q ≥ 0 mass preserving, which can easily be seen by considering the Fourier transform uˆ(k, t) = exp(t[(λ ± ik)α − λα ]) of point source solutions to the tempered fractional diffusion equation (3.23). Now x → u(x, t) are the probability density functions of a tempered stable L´evy process, as in Rosi´ nski [59]. That process arises as the long-time scaling limit of a random walk with exponentially tempered power law jumps, see Chakrabarty and Meerschaert [13]. The tempered fractional diffusion equation (3.23) has been applied to contaminant plumes in underground aquifers, and sediment transport in rivers [49, 70, 71]. Remark 3.1.17. Tempered fractional derivatives are a natural analogues of integer (and fractional) order derivatives. For suitable functions f (x), the Fourier transform of the derivative f ′ (x) is (ik)fˆ(k) (e.g., see [48, p. 8]), and one can define the fractional derivative Dα ± f (t) as the function with Fourier transform (ik)α fˆ(k). Definition 3.1.11 extends to tempered fractional derivatives. 28 3.2 Stochastic Integrals In this section, we apply tempered fractional calculus to define stochastic integrals with respect to TFBM. First we recall the moving average representation of TFBM as a stochastic integral with respect to Brownian motion. Recall from Definition 2.1.1, the stochastic integral ∫ +∞ [ Bα,λ (t) = −∞ −λ(−x)+ (−x)−α e−λ(t−x)+ (t − x)−α + −e + ] B(dx). (3.24) When λ = 0 and α < −1/2, the right-hand side of (3.24) does not exist, since the integrand is not in L2 (R). However, TFBM with λ > 0 and α < −1/2 is well-defined, because the exponential tempering keeps the integrand in L2 (R). In fact, if α < −1/2 and λ > 0, or if α = 0 and λ > 0, we will now show that TFBM is a semimartingale, and hence one ∫ can define stochastic integrals I(f ) := f (x)Bα,λ (dx) in the standard manner, via the Itˆo stochastic calculus (e.g., see Kallenberg [30, Chapter 15]). Theorem 3.2.1. A tempered fractional Brownian motion {Bα,λ (t)}t≥0 with α < −1/2 and λ > 0 is a continuous semimartingale with the canonical decomposition Bα,λ (t) = −λ where ∫ t 0 Mα,λ (s) ds − α ∫ +∞ Mα,λ (t) := −∞ ∫ t 0 Mα+1,λ (s) ds e−λ(t−x)+ (t − x)−α + B(dx). (3.25) (3.26) Moreover, {Bα,λ (t)}t≥0 is a finite variation process. The same is true if α = 0 and λ > 0. Proof. Let {FtB }t≥0 be the σ-algebra generated by {Bs : 0 ≤ s ≤ t}. Given a function 29 g : R → R such that g(t) = 0 for all t < 0, and ∫ t g(t) = C + h(s), ds for all t > 0, (3.27) 0 for some C ∈ R and some h ∈ L2 (R), a result of Cheridito [14, Theorem 3.9] shows that the Gaussian stationary increment process g Yt := ∫ R [g(t − u) − g(−u)] B(du), t ≥ 0 (3.28) is a continuous {FtB }t≥0 semimartingale with canonical decomposition g Yt = g(0)Bt + ∫ t∫ s 0 −∞ h(s − u)B(du)ds, (3.29) and conversely, that if (3.28) defines a semimartingale on [0, T ] for some T > 0, then g satisfies these properties. Define g(t) = 0 for t ≤ 0 and g(t) := e−λt t−α for t > 0. (3.30) It is easy to check that the function g(t − u) − g(−u), which is the integrand in (3.24), is square integrable over the entire real line for any α < 1/2 and λ > 0. Next observe that (3.27) holds with C = 0, h(s) = 0 for s < 0 and h(s) := d −λs −α [e s ] = −λe−λs s−α − αe−λs s−α−1 ∈ L2 (R) ds (3.31) for any α < −1/2 and λ > 0. Then it follows from [14, Theorem 3.9] that TFBM is a 30 continuous semimartingale with canonical decomposition ∫ t∫ s Bα,λ = 0 −∞ −λe−λ(s−u) (s − u)−α − αe−λ(s−u) (s − u)−α−1 B(du) ds (3.32) which reduces to (3.25). Since C = 0, Theorem 3.9 in [14] implies that {Bα,λ (t)} is a finite variation process. The proof for α = 0 is similar, using g(t) = e−λt for t > 0. Remark 3.2.2. When α = 0 and λ > 0, the Gaussian stochastic process (3.26) is an Ornstein-Uhlenbeck process. When α < −1/2 and λ > 0, it is a one dimensional Mat´ern stochastic process [5, 23, 26], also called a “fractional Ornstein-Uhlenbeck process” in the physics literature [40]. It follows from Knight [32, Theorem 6.5] that Mα,λ (t) is a semimartingale in both cases. Cheridito [14, Theorem 3.9] provides a necessary and sufficient condition for the process (3.28) to be a semimartingale, and then it is not hard to check that TFBM is not a semimartingale in the remaining cases when −1/2 < α < 0 or 0 < α < 1/2. Next we will investigate the problem of stochastic integration with deterministic integrands in these two cases. Our approach follows that of Pipiras and Taqqu [56]. Next we establish a link between TFBM and tempered fractional calculus. Lemma 3.2.3. For a tempered fractional Brownian motion (3.24) with λ > 0, we have: (i) When −1/2 < α < 0, we can write ∫ +∞ [ ] κ,λ κ+1,λ Bα,λ (t) = Γ(κ + 1) I− 1[0,t] (x) − λI− 1[0,t] (x) B(dx) −∞ where κ = −α. 31 (3.33) (ii) When 0 < α < 1/2, we can write ∫ +∞ [ ] 1−α,λ α,λ 1[0,t] (x) B(dx). Bα,λ (t) = Γ(1 − α) D− 1[0,t] (x) − λI− −∞ (3.34) Proof. To prove part (i), write the kernel function from (3.24) in the form −λ(−x)+ (−x)−α gt,λ (x) : = e−λ(t−x)+ (t − x)−α + −e + [ ] ∫ t d e−λ(u−x)+ (u − x)κ + = du du 0 ∫ +∞ (κ+1)−1 = −λ 1[0,t] (u)e−λ(u−x)+ (u − x)+ du −∞ ∫ +∞ +κ 1[0,t] (u)e−λ(u−x)+ (u − x)κ−1 + du −∞ and apply the definition (3.2) of the tempered fractional integral. To prove part (ii), it suffices to show that the integrand −λ(0−x)+ (0 − x)−α =: ϕ (x) − ϕ (x) gt,λ (x) = e−λ(t−x)+ (t − x)−α t 0 + −e + in (3.24) equals the integrand in (3.34). We will prove this using Fourier transforms. The substitution u = t − x shows that ∫ t 1 e−ikt Γ(1 − α) , ϕt (k) = √ e−ikx e−λ(t−x) (t − x)−α dx = √ 2π −∞ 2π(λ − ik)1−α using the formula for the Fourier transform of the gamma density, and hence e−ikt − 1 gt,λ (k) = ϕt (k) − ϕ0 (k) = Γ(1 − α) √ . 2π(λ − ik)1−α 32 (3.35) On the other hand, from Lemma 3.1.6 and Theorem 3.1.9 we obtain α,λ 1−α,λ F[D− 1[0,t] − λI− 1[0,t] ](k) = [(λ − ik)α − λ(λ − ik)α−1 ] · e−ikt − 1 √ (−ik) 2π e−ikt − 1 = (λ − ik)α−1 · √ , 2π (3.36) where we have used the formula (which is easy to verify) e−ikb − e−ika ˆ √ h(k) = F[1[a,b) ](k) = . (−ik) 2π (3.37) The desired result now follows by the uniqueness of the Fourier transform. Next, we explain the connection between the fractional calculus representations (3.33) and (3.34). Substitute κ = −α into (3.33) and note that the resulting formula differs from (3.34) −α,λ only in that the tempered fractional integral I− α,λ α,λ is replaced by the tempered fractional α,λ derivative D− . Lemma 3.1.14 shows that I− and D− are inverse operators, and hence it −α,λ makes sense to define I± α,λ := D± when 0 < α < 1. Now equations (3.33) and (3.34) are equivalent. Next, we discuss a general construction for stochastic integrals with respect to TFBM. For a standard Brownian motion {B(t)}t∈R on (Ω, F, P ), the stochastic integral I(f ) := ∫ f (x)B(dx) is defined for any f ∈ L2 (R), and the mapping f → I(f ) defines an isometry from L2 (R) into L2 (Ω), called the Itˆ o isometry: ∫ ⟨I(f ), I(g)⟩L2 (Ω) = Cov[I(f ), I(g)] = f (x)g(x) dx = ⟨f, g⟩L2 (R) . (3.38) Since this isometry maps L2 (R) onto the space Sp(B) = {I(f ) : f ∈ L2 (R)}, we say that 33 these two spaces are isometric. For any elementary function (step function) f (u) = n ∑ i=1 ai 1[t ,t ) (u), i i+1 (3.39) where ai , ti are real numbers such that ti < tj for i < j, it is natural to define the stochastic integral ∫ I α,λ (f ) = R f (x)Bα,λ (dx) = n ∑ [ ] ai Bα,λ (ti+1 ) − Bα,λ (ti ) , (3.40) i=1 and then it follows immediately from (3.33) that for f ∈ E, the space of elementary functions, the stochastic integral ∫ I α,λ (f ) = R ∫ [ ] κ,λ κ+1,λ f (x)Bα,λ (dx) = Γ(κ + 1) I− f (x) − λI− f (x) B(dx) R is a Gaussian random variable with mean zero, such that for any f, g ∈ E we have (∫ ⟨I α,λ (f ), I α,λ (g)⟩ = Γ(κ + 1)2 ∫ [ R =E L2 (Ω) R ∫ f (x)Bα,λ (dx) R ) g(x)Bα,λ (dx) ][ ] κ,λ κ+1,λ κ,λ κ+1,λ I− f (x) − λI− f (x) I− g(x) − λI− g(x) dx, (3.41) in view of (3.33) and the Itˆo isometry (3.38). The linear space of Gaussian random variables { } I α,λ (f ), f ∈ E is contained in the larger linear space { } Sp(Bα,λ ) = X : I α,λ (fn ) → X in L2 (Ω) for some sequence (fn ) in E . An element X ∈ Sp(Bα,λ ) is mean zero Gaussian with variance Var(X) = lim Var[I α,λ (fn )], n→∞ 34 (3.42) and X can be associated with an equivalence class of sequences of elementary functions (fn ) such that I α,λ (fn ) → X in L2 (R). If [fX ] denotes this class, then X can be written in an integral form as ∫ X= R [fX ]dBα,λ (3.43) and the right hand side of (3.43) is called the stochastic integral with respect to TFBM on the real line (see, for example, Huang and Cambanis [27], page 587). In the special case of a Brownian motion λ = α = 0, I α,λ (fn ) → X along with the Itˆo isometry (3.38) implies that (fn ) is a Cauchy sequence, and then since L2 (R) is a (complete) Hilbert space, there exists a unique f ∈ L2 (R) such that fn → f in L2 (R), and we can write X = ∫ R f (x)B(dx). However, if the space of integrands is not complete, then the situation is more complicated. We begin with the case −1/2 < α < 0, where the corresponding FBM is long range dependent. 3.2.1 Semi-long range dependence Here we investigate stochastic integrals with respect to TFBM in the case −1/2 < α < 0, so that 1/2 < H < 1 in (2.4). Equation (3.41) suggests the appropriate space of integrands for TFBM, in order to obtain a nice isometry that maps into the space Sp(Bα,λ ) of stochastic integrals. Theorem 3.2.4. Given −1/2 < α < 0 and λ > 0, let κ = −α. Then the class of functions { ∫ 2 A1 := f ∈ L (R) : R } 2 κ,λ κ+1,λ I− f (x) − λI− f (x) dx < ∞ 35 , (3.44) is a linear space with inner product ⟨f, g⟩A := ⟨F, G⟩L2 (R) (3.45) 1 where κ+1,λ κ,λ F (x) = Γ(κ + 1)[I− f (x) − λI− f (x)], (3.46) κ,λ κ+1,λ G(x) = Γ(κ + 1)[I− g(x) − λI− g(x)]. The set of elementary functions E is dense in the space A1 . The space A1 is not complete. κ,λ κ+1,λ The proof of Theorem 3.2.4 requires one simple lemma, which shows that I− − λI− is a bounded linear operator on Lp (R) for any 1 ≤ p < ∞. Lemma 3.2.5. Under the assumptions of Theorem 3.2.4, suppose 1 ≤ p < ∞. Then for any f ∈ Lp (R) we have κ,λ κ+1,λ ∥I− f (x) − λI− f (x)∥p ≤ C∥f ∥p (3.47) where C is a constant depending only on α and λ. κ,λ κ+1,λ Proof. It follows from Lemma 3.1.2 that I− f (x) − λI− κ,λ κ+1,λ ∥I− f (x) − λI− κ,λ f (x) ∈ Lp (R) and that κ+1,λ f (x)∥p ≤ ∥I− f (x)∥p + λ∥I− f (x)∥p ≤ 2λ−κ ∥f ∥p for any f ∈ Lp (R). Remark 3.2.6. It follows from Lemma 3.2.5 that A1 contains every function in L2 (R), and hence they are the same set, but endowed with a different inner product. The inner product 36 on the space A1 is required to obtain a nice isometry. Proof of Theorem 3.2.4. The proof is similar to [56, Theorem 3.2]. To show that A1 is an inner product space, we will check that ⟨f, f ⟩A = 0 implies f = 0 almost everywhere. If 1 ⟨f, f ⟩A = 0, then in view of (3.45) and (3.46) we have ⟨F, F ⟩2 = 0, so F (x) = Γ(1 + 1 κ,λ κ+1,λ κ)[I− f (x) − λI− f (x)] = 0 for almost every x ∈ R. Then κ,λ κ+1,λ I− f (x) = λI− f (x) for almost every x ∈ R. (3.48) κ,λ Apply D− to both sides of equation (3.48) and use Lemma 3.1.4 along with Lemma 3.1.14 to get κ,λ κ,λ κ,λ κ+1,λ f (x) = D− I− f (x) = D− λI− [ ] κ,λ κ,λ 1,λ 1,λ f (x) = λ D− I− I− f (x) = λI− f (x) for almost every x ∈ R, and in view of the definition (3.1) this is equivalent to ∫ +∞ f (x) = λ f (u)e−λ(u−x) du = λeλx x ∫ +∞ f (u)e−λu du (3.49) x for almost every x ∈ R. Observe that the functions f (u) and e−λu are in L2 [x, ∞) for any x ∈ R and then, by the Cauchy-Schwartz inequality, the function f (u)e−λu is in L1 [x, ∞). It ∫ follows that x+∞ f (u)e−λu du is absolutely continuous, and so the function f (x) in (3.49) is also absolutely continuous. Taking the derivative on both sides of (3.49) using the Lebesgue Differentiation Theorem (e.g., see [66, Theorem 7.16]) we get f ′ (x) = λf (x) − λeλx f (x)e−λx = 0 for almost every x ∈ R. 37 Then for any a, b ∈ R we have ∫ b f (b) = f (a) + f ′ (x) dx = f (a). a and so f (x) is a constant function. Since f ∈ L2 (R), it follows that f (x) = 0 for all x ∈ R, and hence A1 is an inner product space. Next, we want to show that the set of elementary functions E is dense in A1 . For any f ∈ A1 , we also have f ∈ L2 (R), and hence there exists a sequence of elementary functions (fn ) in L2 (R) such that ∥f − fn ∥2 → 0. But ∥f − fn ∥A1 = ⟨f − fn , f − fn ⟩A = ⟨F − Fn , F − Fn ⟩2 = ∥F − Fn ∥2 , 1 κ,λ κ+1,λ where Fn (x) = I− fn (x) − λI− fn (x) and F (x) is given by (3.46). Lemma 3.2.5 implies that κ,λ κ+1,λ ∥f − fn ∥A1 = ∥F − Fn ∥2 = ∥I− (f − fn ) − λI− (f − fn )∥2 ≤ C∥f − fn ∥2 for some C > 0, and since ∥f − fn ∥2 → 0, it follows that the set of elementary functions is dense in A1 . Finally, we provide an example to show that A1 is not complete. The functions fn (k) = |k|−p 1{1<|k| 0, are in L2 (R), fn (k) = fn (−k), and hence they are the Fourier transforms of functions fn ∈ L2 (R). Apply Lemma 3.1.6 to see that the corresponding functions Fn (x) = Γ(κ + 38 κ,λ κ+1,λ 1)[I− fn (x) − λI− fn (x)] from (3.46) have Fourier transform −ikΓ(1 − α) ˆ F[Fn ](k) = Γ(1 − α)[(λ − ik)α − λ(λ − ik)α−1 ]fˆn (k) = fn (k). (λ − ik)1−α (3.50) Since α < 0, it follows that ∥Fn ∥22 = ∥Fˆn ∥22 = Γ(1 − α)2 ∫ ∞ −∞ fn (k) 2 k2 dk < ∞ (λ2 + k 2 )1−α for each n, which shows that fn ∈ A1 . Now it is easy to check that fn − fm → 0 in A1 , as n, m → ∞, whenever p > 1/2 + α, so that (fn ) is a Cauchy sequence. Choose p ∈ (1/2 + α, 1/2) and suppose that there exists some f ∈ A1 such that ∥fn − f ∥A1 → 0 as n → ∞. Then ∫ ∞ −∞ fn (k) − f (k) 2 k2 dk → 0 (λ2 + k 2 )1−α (3.51) as n → ∞, and since, for any given m ≥ 1, the value of fn (k) does not vary with n > m whenever k ∈ [−m, m], it follows that fˆ(k) = |k|−p 1{|k|>1} on any such interval. Since m is arbitrary, it follows that fˆ(k) = |k|−p 1{|k|>1} , but this function is not in L2 (R), so fˆ(k) ∈ / A1 , which is a contradiction. Hence A1 is not complete, and this completes the proof. We now define the stochastic integral with respect to TFBM for any function in A1 in the case where 1/2 < H < 1 in (2.4). Definition 3.2.7. For any −1/2 < α < 0 and λ > 0, we define ∫ R ∫ [ f (x)Bα,λ (dx) := Γ(κ + 1) R κ,λ κ+1,λ I− f (x) − λI− 39 ] f (x) B(dx) (3.52) for any f ∈ A1 , where κ = −α. Theorem 3.2.8. For any −1/2 < α < 0 and λ > 0, the stochastic integral I α,λ in (3.52) is an isometry from A1 into Sp(Bα,λ ). Since A1 is not complete, these two spaces are not isometric. Proof. It follows from Lemma 3.2.5 that the stochastic integral (3.52) is well-defined for any f ∈ A1 . Proposition 2.1 in Pipiras and Taqqu [56] implies that, if D is an inner product space such that (f, g)D = ⟨I α,λ (f ), I α,λ (g)⟩L2 (Ω) for all f, g ∈ E, and if E is dense D, then there is an isometry between D and a linear subspace of Sp(Bα,λ ) that extends the map f → I α,λ (f ) for f ∈ E, and furthermore, D is isometric to Sp(Bα,λ ) itself if and only if D is complete. Using the Itˆo isometry and the definition (3.52), it follows from (3.45) that for any f, g ∈ A1 we have ⟨f, g⟩A = ⟨F, G⟩L2 (R) = ⟨I α,λ (f ), I α,λ (g)⟩L2 (Ω) , 1 and then the result follows from Theorem 3.2.4. 3.2.2 Anti-persistence Next we investigate stochastic integrals with respect to TFBM in the case 0 < α < 1/2, so that 0 < H < 1/2 in (2.4). It follows from (3.34) that the stochastic integral (3.40) can be written in the form ∫ I α,λ (f ) = R f (x)Bα,λ (dx) = Γ(1 − α) ∫ ∞ [ 40 −∞ α,λ 1−α,λ D− f (x) − λI− ] f (x) B(dx) for any f ∈ E, the space of elementary functions. Then I α,λ (f ) is a Gaussian random variable with mean zero, such that (∫ ⟨I α,λ (f ), I α,λ (g)⟩ =E L2 (Ω) R ∫ f (x)Bα,λ (dx) R ) g(x)Bα,λ (dx) ∫ [ ][ ] α,λ 1−α,λ α,λ 1−α,λ 2 D− f (x) − λI− f (x) D− g(x) − λI− g(x) dx. = Γ(1 − α) (3.53) R for any f, g ∈ E, using (3.34) and the Itˆo isometry (3.38). Equation (3.53) suggests the following space of integrands for TFBM in the case 0 < H < 1/2. Recall that W α,2 (R) is the fractional Sobolev space (3.17). Theorem 3.2.9. For any 0 < α < 1/2 and λ > 0, the class of functions { } α,λ 1−α,λ α,2 2 A2 := f ∈ W (R) : φf = D− f − λI− f for some φf ∈ L (R). . (3.54) is a linear space with inner product ⟨f, g⟩A := ⟨F, G⟩L2 (R) (3.55) [ ] α,λ 1−α,λ F (x) = Γ(1 − α) D− f (x) − λI− f (x) [ ] α,λ 1−α,λ G(x) = Γ(1 − α) D− g(x) − λI− g(x) . (3.56) 2 where The set of elementary functions E is dense in the space A2 . The space A2 is not complete. We begin with the two lemmas. The first lemma shows that the set A2 contains every function in W α,2 (R), and hence they are the same set, but different spaces, since they have 41 different inner products. Lemma 3.2.10. Under the assumptions of Theorem 3.2.9, every f ∈ W α,2 (R) is an element of A2 . Proof. Given f ∈ W α,2 (R), we need to show that 1−α,λ α,λ φf = D− f − λI− f, for some φf ∈ L2 (R). From the definition (3.17) we see that (3.57) ∫ (λ2 + k 2 )α |fˆ(k)|2 dk < ∞. Define h1 (k) = (λ − ik)α fˆ(k) and note that h1 is the Fourier transform of some function φ1 ∈ L2 (R). Define h2 (k) := (λ − ik)α−1 f (k), and observe that ∫ ∫ |h2 (k)|2 dk = |f (k)|2 (λ2 + k 2 )α−1 dk ∫ = |h1 (k)|2 dk < ∞, λ2 + k 2 since h1 ∈ L2 (R) and 1/(λ2 + k 2 ) is bounded. Hence there is another function φ2 ∈ L2 (R) such that h2 = φˆ2 . Define φf := φ1 − λφ2 so that φf (k) = φ1 (k) − λφ2 (k) = f (k)(λ − ik)α − f (k)λ(λ − ik)α−1 . (3.58) Since f ∈ W α,2 (R) ⊂ L2 (R), we can apply Definition 3.1.11 and Lemma 3.1.6 to see that (3.57) holds. Lemma 3.2.11. Under the assumptions of Theorem 3.2.9, if f ∈ W α,2 (R), then there exists 42 a sequence of elementary functions (fn ) such that fn → f in L2 (R), and also ∫ +∞ −∞ |fn (k) − f (k)|2 |k|2α dk → 0 as n → ∞. (3.59) Proof. Equation (3.59) is proven in [56, Lemma 5.1]. For any L > 0, that proof constructs a sequence of elementary functions fn such that fˆn (k) → 1[−1,1] (k) almost everywhere on −L ≤ x ≤ L, and shows that |fˆn (k)| ≤ C min{1, |k|−1 } for all k ∈ R and all n ≥ 1. In the notation of that paper, we have fˆn (k) = k −1 Un (k). Apply the dominated convergence theorem to see that ∫ +L −L and note that |fˆn (k) − 1[−1,1] (k)|2 dk → 0 ∫ |k|>L |fˆn (k) − 1[−1,1] (k)|2 dk ≤ 2C 2 ∫ ∞ dk 2C 2 ≤ . 2 L L k Since L is arbitrary, it follows that fˆn (k) → 1[−1,1] (k) in L2 (R), and then the result follows as in [56, Lemma 5.1]. Proof of Theorem 3.2.9. For f ∈ A2 we define ∥f ∥A2 = √ ⟨f, f ⟩A = √ 2 ⟨φf , φf ⟩2 = ∥φf ∥2 . (3.60) where φf is given by (3.57). Next, use (3.58) to see that φf (k) = (−ik)(λ − ik)α−1 f (k). 43 (3.61) To verify that (3.55) is an inner product, note that if ⟨f, f ⟩A = 0 then 2 ∥f ∥2A = ∥φf ∥22 = ∥φf ∥22 = 2 ∫ ∞ −∞ |f (k)|2 k2 dk (λ2 + k 2 )1−α (3.62) equals zero, which implies that |f (k)| = 0 almost everywhere, and then f = 0 almost everywhere. This proves that (3.57) is an inner product. Next we show that E is dense in A2 . Apply Lemma 3.2.11 to obtain a sequence (fn ) in E such that ∥fn − f ∥2 → 0 and (3.59) holds. It is easy to check using (3.37) that any elementary function is an element of W α,2 (R), and then Lemma 3.2.10 implies that it is also an element of A2 . Now use (3.62) to write ∥fn − f ∥2A = 2 ∫ +∞ −∞ − λ2 2 fn (k) − f (k) (k 2 + λ2 )α dk ∫ +∞ −∞ fn (k) − f (k) 2 1 (λ2 + k 2 )1−α dk. Since 1/(λ2 + k 2 )1−α is bounded, it follows easily using (3.59) and ∥fn − f ∥2 → 0 that ∥fn − f ∥A2 → 0, and hence E is dense in A2 . Finally, we want to show that A2 is not complete. The proof is similar to that of Theorem 3.2.4. The functions fn (k) = |k|−p 1{1/n<|k|<1} (k). are the Fourier transforms of some functions fn ∈ L2 (R). Clearly fn ∈ W α,2 (R), and then it follows from Lemmas 3.1.6 and 3.1.9 that the corresponding functions Fn (x) = α,λ 1−α,λ Γ(1 − α)[D− fn (x) − λI− fn (x)] from (3.56) have Fourier transform (3.50), that is, F[Fn ](k) = Γ(1 − α) 44 −ik fˆn (k). (λ − ik)1−α Then ∫ ∞ 2 k2 2 2 2 2 ∥fn ∥A = ∥Fn ∥2 = ∥Fn ∥2 = Γ(1 − α) fn (k) dk < ∞ 2 (λ2 + k 2 )1−α −∞ for any p < 3/2, so that fn ∈ A2 . Now it is easy to check that fn − fm → 0 in A2 , as n, m → ∞, so that (fn ) is a Cauchy sequence. Suppose 1/2 < p < 3/2 and that ∥fn − f ∥A2 → 0 for some f ∈ A2 . Then fˆ(k) = |k|−p 1{0<|k|<1} , but this fˆ is not in L2 (R), so fˆ ∈ / A2 , and hence A2 is not complete. We now define the stochastic integral with respect to TFBM for any function in A2 in the case where 0 < H < 1/2 in (2.4). Definition 3.2.12. For any 0 < α < 1/2 and λ > 0, we define ∫ I α,λ (f ) = R ∫ [ ] α,λ 1−α,λ f (x)Bα,λ (dx) : = Γ(1 − α) D− f (x) − λI− f (x) B(dx) (3.63) R for any f ∈ A2 . Theorem 3.2.13. For any 0 < α < 1/2 and λ > 0, the stochastic integral I α,λ is an isometry from A2 into Sp(Bα,λ ). Since A2 is not complete, these two spaces are not isometric. Proof. The proof is similar to that of Theorem 3.2.8. It follows from Lemma 3.2.10 that the stochastic integral (3.63) is well-defined for any f ∈ A2 . Use Proposition 2.1 in Pipiras and Taqqu [56], and note that the Itˆo isometry, the definition (3.63), and equation (3.55) imply that for any f, g ∈ A2 we have ⟨f, g⟩A = ⟨F, G⟩L2 (R) = ⟨I α,λ (f ), I α,λ (g)⟩L2 (Ω) . 2 45 Then the result follows from Theorem 3.2.9. 3.2.3 Harmonizable representation By now it should be clear that the Fourier transform plays an important role in the theory of stochastic integration for TFBM. Here we apply the harmonizable representation of TFBM to unify the two cases −1/2 < α < 0 and 0 < α < 1/2. For any −1/2 < α < 1/2 and any λ > 0, Proposition 3.1 in [46] shows that TFBM has the harmonizable representation ∫ Γ(1 − α) ∞ e−itk − 1 ˆ B(dk) Bα,λ (t) = √ 1−α 2π −∞ (λ − ik) ˆ = B ˆ1 + iB ˆ2 is a complex-valued Gaussian random measure constructed as folwhere B ˆ1 and B ˆ2 be two independent Brownian motions on the positive real line with lows. Let B ˆi (t))2 ] = t/2 for i = 1, 2, and define two independently scattered Gaussian random E[(B ˆi [a, b] = B ˆi (b) − B ˆi (a), extend to Borel subsets of the positive real line, measures by setting B ˆ1 (A) = B ˆ1 (−A), B ˆ2 (A) = −B ˆ2 (−A). and then extend to the entire real line by setting B Apply the formula (3.37) for the Fourier transform of an indicator function to write this harmonizable representation in the form Bα,λ (t) = Γ(1 − α) ∫ +∞ −∞ 1[0,t] (k) (−ik) ˆ B(dk). (λ − ik)1−α It follows easily that for any elementary function (3.39) we may write I α,λ (f ) = Γ(1 − α) ∫ ∞ −∞ f (k) 46 (−ik) ˆ B(dk), (λ − ik)1−α (3.64) and then for any elementary functions f and g we have ⟨I α,λ (f ), I α,λ (g)⟩ L2 (Ω) = Γ(1 − α)2 ∫ ∞ −∞ f (k)g(k) k2 dk. (λ2 + k 2 )1−α (3.65) Theorem 3.2.14. For any α ∈ (−1/2, 0) ∪ (0, 1/2) and λ > 0, the class of functions { ∫ 2 2 A3 := f ∈ L (R) : f (k) } k2 dk < ∞ . (λ2 + k 2 )1−α (3.66) is a linear space with the inner product ∫ +∞ k2 2 f (k)g(k) 2 ⟨f, g⟩A = Γ(1 − α) dk. 3 (λ + k 2 )1−α −∞ (3.67) The set of elementary functions E is dense in the space A3 . The space A3 is not complete. Proof. The proof combines Theorems 3.2.4 and 3.2.9 using the Plancherel Theorem. First α,λ 1−α,λ suppose that 0 < α < 1/2 and recall that φf = D− f − λI− f is a function with Fourier transform φˆf = [(λ − ik)α − λ(λ − ik)α−1 ]fˆ = [λ − ik − λ](λ − ik)α−1 fˆ = (−ik)(λ − ik)α−1 fˆ. Then it follows from the Plancherel Theorem that ⟨f, g⟩A = Γ(1 − α)2 ⟨φf , φg ⟩2 = Γ(1 − α)2 ⟨φˆf , φˆg ⟩2 2 ∫ +∞ 2 = Γ(1 − α) f (k)g(k) −∞ 47 k2 dk = ⟨f, g⟩A 3 (λ2 + k 2 )1−α and hence the two inner products are identical. If f ∈ A3 , then ∫ +∞ −∞ |fˆ(k)|2 (λ2 + k 2 )α dk = ∫ +∞ −∞ + λ2 |fˆ(k)|2 ∫ +∞ −∞ k2 dk (λ2 + k 2 )1−α |fˆ(k)|2 1 dk. (λ2 + k 2 )1−α (3.68) The first integral on the right-hand side is finite by (3.66), and the second is finite since 1/(λ2 + k 2 )1−α is bounded. Then it follows from the definition (3.17) that f ∈ W α,2 (R). Conversely, if f ∈ W α,2 (R) then since k2 k2 = (λ2 + k 2 )α ≤ (λ2 + k 2 )α (λ2 + k 2 )1−α λ2 + k 2 it follows immediately that f ∈ A3 , and hence W α,2 (R) and A3 are the same set of functions. Then it follows from Lemma 3.2.10 that A2 and A3 are identical when 0 < α < 1/2, and the conclusions of Theorem 3.2.14 follow from Theorem 3.2.9 in this case. If −1/2 < α < 0, then the function k 2 /(λ2 + k 2 )1−α is bounded by a constant C(α, λ) that depends only on α and λ, so for any f ∈ L2 (R) we have ∫ R f (k) 2 ∫ 2 k2 dk ≤ C(α, λ) f (k) dk < ∞ (λ2 + k 2 )1−α R (3.69) and hence f ∈ A3 . Since A3 ⊂ L2 (R) by definition, this proves that L2 (R) and A3 are the same set of functions, and then it follows from Lemma 3.2.5 that A1 and A3 are the same κ,λ κ+1,λ set of functions in this case. Let κ = −α and note that φf = I− f − λI− function with Fourier transform φˆf = [(λ − ik)α − λ(λ − ik)α−1 ]fˆ = (−ik)(λ − ik)α−1 fˆ. 48 f is again a Then it follows from the Plancherel Theorem that ⟨f, g⟩A = Γ(κ + 1)2 ⟨φf , φg ⟩2 = Γ(1 − α)2 ⟨φˆf , φˆg ⟩2 1 ∫ +∞ 2 = Γ(1 − α) f (k)g(k) −∞ k2 dk = ⟨f, g⟩A 3 (λ2 + k 2 )1−α and hence the two inner products are identical. Then the conclusions of Theorem 3.2.14 follow from Theorem 3.2.4 in this case as well. Definition 3.2.15. For any α ∈ (−1/2, 0) ∪ (0, 1/2) and λ > 0, we define I α,λ (f ) = Γ(1 − α) ∫ ∞ −∞ f (k) (−ik) ˆ B(dk) (λ − ik)1−α (3.70) for any f ∈ A3 . Theorem 3.2.16. For any α ∈ (−1/2, 0) ∪ (0, 1/2) and λ > 0, the stochastic integral I α,λ in (3.70) is an isometry from A3 into Sp(Bα,λ ). Since A3 is not complete, these two spaces are not isometric. Proof. The proof of Theorem 3.2.14 shows that A1 and A3 are identical when −1/2 < α < 0, and A2 and A3 are identical when 0 < α < 1/2. Then the result follows immediately from Theorems 3.2.8 and 3.2.13. 3.3 Discussion In this section, we collect some remarks and extensions. 49 3.3.1 White noise approach Heuristically, the TFBM (3.33) with 1/2 < H < 1 in (2.4) can be written in terms of tempered fractional integrals of the white noise W (x)dx = B(dx), since in view of (3.8) we can write ∫ +∞ [ ] κ+1,λ κ,λ W (x) 1[0,t] (x) dx. Bα,λ (t) = Γ(κ + 1) I+ W (x) − λI+ −∞ In the same way, when 0 < H < 1/2 we can write Bα,λ (t) = Γ(1 − α) ∫ +∞ [ −∞ ] α,λ 1−α,λ W (x) 1[0,t] (x) dx, D+ W (x) − λI+ using Lemma 3.1.15. These ideas could be made rigorous using white noise theory [36]. Setting λ = 0, we recover the fact that FBM is the fractional integral or derivative of a Brownian motion [56, p. 261]. The white noise approach is preferred in engineering applications (e.g., see [6]). 3.3.2 Reproducing kernel Hilbert space The reproducing kernel Hilbert space (RKHS) of TFBM provides another approach to stochastic integration that produces an isometric space of deterministic integrands. The RKHS for FBM was computed in [6, 56]. For any mean zero Gaussian process {Xt }t∈R with covariance function R(s, t) = E[Xs Xt ], the RKHS of X is the unique Hilbert space H(X) of measurable functions f : R → R such that R(·, t) ∈ H(X) for all t ∈ R, and ⟨f, R(·, t)⟩H(X) = f (t) for all t ∈ R and f ∈ H(X) [25, 65]. As noted in [25], if there exists 50 a measure space (Λ, B, ν) and a set of functions {ft } ⊂ L2 (R, ν) such that ∫ R(s, t) = Λ fs (x)ft (x)ν(dx) for all s, t ∈ R, Then H(X) consists of the functions g(t) = ∫ (3.71) ft (x)g ∗ (x)ν(dx) for g ∗ ∈ Sp{ft }, the closure in L2 (R, ν) of the set of linear combinations of functions ft . Then H(X) is a Hilbert space with the inner product ∫ ⟨g, h⟩H(X) = g ∗ (x)h∗ (x)ν(dx). Λ Let Sp(X) denote the closure of the set of linear combinations of random variables {Xt } in the space L2 (Ω). The mapping J that sends J ∑ aj R(·, tj ) → j=1 J ∑ aj Xtj j=1 is an isometry that maps H(X) onto Sp(X), and hence these two Hilbert spaces are isometric. Then J (f ) is the stochastic integral of any f ∈ H(X). ∫ For TFBM with −1/2 < α < 0, let κ = −α. Since Bα,λ (t) = R 1[0,t] (x)Bα,λ (dx), it follows immediately from the definition (3.52) that TFBM has covariance function R(s, t) = Γ(κ + 1)2 ∫ [ ][ ] κ,λ κ+1,λ κ,λ κ+1,λ I− 1[0,s] (x) − λI− 1[0,s] I− 1[0,t] (x) − λI− 1[0,t] dx, R and hence the RKHS H(Bα,λ ) consists of functions ∫ [ ] k,λ k+1,λ g(t) = Γ(k + 1) I− − λI− 1[0,t] (x)g ∗ (x)dx R 51 for g ∗ ∈ L2 (R), with the inner product ∫ ⟨g, h⟩H(X) = R g ∗ (x)h∗ (x)dx = ⟨g ∗ , h∗ ⟩L2 (R) . (3.72) For TFBM with 0 < α < 1/2 and λ > 0, the RKHS H(Bα,λ ) consists of functions g(t) = Γ(1 − α)2 ∫ [ ] 1−α,λ α,λ 1[0,t] (x)g ∗ (x)dx D− − λI− R for g ∗ ∈ L2 (R), with the same inner product (3.72). The proof is similar to [56, Section 6]. Complete details will be provided in the forthcoming paper [45]. Here we take Λ = L2 (R), with ν the Lebesgue measure on R. The main technical difficulty is to show that k,λ k+1,λ L2 (R) = Sp{ft }, where ft (x) = Γ(k + 1)[I− − λI− α,λ 1−α,λ and ft (x) = Γ(1 − α)[D− − λI− 3.3.3 ]1[0,t] (x) in the case −1/2 < α < 0, ]1[0,t] (x) for 0 < α < 1/2. Tempered distributions as integrands Jolis [29] proved that the exact domain of the Wiener integral for a fractional Brownian motion BH (t) is given by ΛH = {f ∈ S ′ (R) = ∫ R |f (k)|2 |k|1−2H dk < ∞}, where S ′ (R) is the space of tempered distributions. This gives an isometry using the inner product (for a standard FBM) Γ(2H + 1) sin(πH) ⟨f, g⟩ = 2π 52 ∫ fˆ(k)ˆ g (k)|k|1−2H dk, that makes ΛH isometric to Sp(BH ). She also proved that this space contains distributions that cannot be represented by locally integrable functions in the case of long range dependence (1/2 < H < 1). Tudor [63] extended this result to subfractional Brownian motion. The distributional approach is useful in the study of partial differential equations with a Gaussian forcing term [9, 15, 64]. Following along these lines, we conjecture that the exact domain of the Wiener integral with respect to TFBM is given by the distributional fractional Sobolev space Λα,λ = {f ∈ S ′ (R) : ∫ R |f (k)|2 (λ2 + k 2 )α dk < ∞} with the inner product ∫ ⟨f, g⟩ = Cα,λ fˆ(k)ˆ g (k)(λ2 + k 2 )α dk. Proving this using [29, Theorem 3.5] would require computing the second derivative of the variance function (2.5) and taking the (inverse) Fourier transform of the result. This computation seems difficult, due to the Bessel function term. 53 Chapter 4 Tempered fractional stable motion This chapter has four sections. In Section 4.1, we define linear tempered fractional stable motion (LTFSM) using a moving average representation, and we establish the dependence structure of its increments, which we call tempered fractional stable noise (TFSN). Section 4.2 defines tempered fractional harmonizable stable motion (HTFSM) and shows that LTFSM is different from HTFSM. Sample path properties of LTFSM and TFHSM are proven in Section 4.3. Finally, Section 4.4 investigates the local times and local nondeterminism properties for LTFSM and HTFSM . 4.1 Moving average representation Let X be a real-valued random variable. We say that X has a symmetric α-stable, SαS, distribution if its characteristic function has the form E [exp {i(θX)}] = exp {−c|θ|α } , for some constant c > 0 and 0 < α ≤ 2. The parameters α and σ are called the index of stability and the scale parameter respectively (see Chapter 1 in [61]). We denote the SαS distribution by Sα (σ, 0, 0) and write X ≃ Sα (σ, 0, 0) to indicate that that X has the stable distribution Sα (σ, 0, 0). A real-valued stochastic process {X(t)} is called SαS if all linear 54 combinations ∑n j=1 θj X(tj ) are real-valued SαS random variables. Consider (R, B(R), dx), where B(R) is the σ-algebra of Borel subsets of R. We say the process {Zα (B), B ∈ B(R)} is a real-valued SαS random measure with Lebesgue control measure dx if [ { }] E exp iRe(θZα (B)) = exp {−|B||θ|α } , for any θ ∈ C where |B| denotes the Lebesgue measure of B (B ∈ B(R)). Now, Let Zα be a SαS random measure on (R, B(R)) with Lebesgue control measure dx. Then, we define the stochastic integral ∫ +∞ I(f ) := −∞ f (x) Zα (dx) (4.1) for all measurable functions f : R → R satisfying the condition ∫ +∞ −∞ f (x) α m(dx) < ∞ (4.2) for any 0 < α < 2, (α ̸= 1). We denote the collection of functions satisfying (4.2) by Lα (R) = ∫ +∞ { } α f : f : R → R is measurable, f (x) dx < ∞ . −∞ (4.3) Proposition 3.4.1 in [61] shows that I(f ), for any f ∈ Lα (R, B(R), dx), has the characteristic function } [ ] { ∫ +∞ α E exp{iθI(f )} = exp − θf (x) dx . −∞ 55 For the stochastic integral I(f ) in (4.1) we define α I(f ) α ∫ +∞ = ( := f (x) Zα (dx) −∞ α [ ]) − log E exp{i I(f )} ∫ +∞ = α −∞ f (x) α (4.4) dx for any 0 < α < 2. Definition 4.1.1. A stochastic process {X(t), t ∈ R} is called an SαS L´evy motion if 1. X(0) = 0 a.s. 2. X has independent increments. 1 3. X(t) − X(s) ∼ Sα (σ α |t − s| α , 0, 0) for any −∞ < s < t < ∞ and, for some 0 < α ≤ 2, and some σ > 0. Note that the process X has stationary increments. It is Brownian motion when α = 2. Also it is α1 - self similar , that is, for all c > 0, { } X(ct) t∈R where { } 1 c α X(t) , t∈R indicates equality in the sense of finite dimensional distributions. Definition 4.1.2. Given a SαS random measure Zα (dx) on R with control measure m(dx), for any 0 < α ≤ 2 and H > 0 and λ ≥ 0, the stochastic integral ∫ +∞ [ XH,α,λ (t) := −∞ 1 1] H− α H− α −λ(t−x) −λ(−x) + + e (t − x)+ −e (−x)+ Zα (dx), 56 (4.5) where (x)+ = max{x, 0} and 00 = 0 will be called a linear tempered fractional stable motion (LTFSM). Remark 4.1.3. When α = 2, {XH,α,λ (t)}t∈R is TFBM defined in (2.1). It is easy to check that the function 1 1 H− H− gα,λ,t (x) := e−λ(t−x)+ (t − x)+ α − e−λ(−x)+ (−x)+ α (4.6) belongs to the linear space Lα (R) in (4.3) provided that H > 0, so that LTFSM is well defined and we have XH,α,λ (t) α α ∫ = ( := ∫ = R gα,λ,t (x) Zα (dx) α α [ ]) − log E exp{i XH,α,λ (t)} R gα,λ,t (x) α (4.7) dx, for any 0 < α < 2. Note also that this function has a scaling property 1 gα,λ,ct (cx) = cH− α gα,cλ,t (x), (4.8) for all t, x ∈ R and all c > 0. The next result shows that LTFSM has a nice scaling property, involving both the time scale and the tempering. Proposition 4.1.4. LTFSM (4.5) is symmetric α-stable stochastic process with stationary increments, such that { } XH,α,λ (ct) t∈R { } cH XH,α,cλ (t) t∈R 57 , (4.9) for any scale factor c > 0. Proof. Since Zα (dx) has control measure m(dx) = σ α dx, the random measure Zα (c dx) has 1 control measure c α σ α dx. Given t1 < t2 < · · · < tn , a change of variable x = cx′ then yields ( (∫ ) XH,α,λ (cti ) : i = 1, . . . , n = (∫ = (∫ ≃ ) gα,λ,cti (x)M (dx) : i = 1, . . . , n gα,λ,cti (cx′ )M (c dx′ ) : i = 1, . . . , n 1 cH− α g ) ′ 1 ′ α,cλ,ti (x )c α M (dx ) : i = 1, . . . , n ) ( ) = cH XH,α,cλ (ti ) : i = 1, . . . , n , where ≃ denotes equality in distribution, so that (4.9) holds. For any s, t ∈ R, the integrand (4.6) satisfies gα,λ,s+t (s + x) − gα,λ,s (s + x) = gα,λ,t (x), and hence a change of variable x = s + x′ yields ( ) XH,α,λ (s + ti ) − XH,α,λ (s) : i = 1, . . . , n ) (∫ [ ] gα,λ,s+ti (x) − gα,λ,s (x) M (dx) : i = 1, . . . , n = ) (∫ [ ] ′ ′ ′ gα,λ,s+ti (s + x ) − gα,λ,s (s + x ) M (dx ) : i = 1, . . . , n ≃ (∫ ) ′ ′ gα,λ,ti (x )M (dx ) : i = 1, . . . , n = ( ) = XH,α,λ (ti ) : i = 1, . . . , n , which shows that LTFSM has stationary increments. When a stochastic process {Y (t)}t∈R is stationary Gaussian with mean zero, we can describe its dependence structure by its covariance function E [Y (t)Y (0)]. However, in the non-Gaussian stable case, the covariance function does not exist. Instead, we use the follow- 58 ing. For a stationary process {Y (t)}, let r(t) : = r(θ1 , θ2 , t) = E [exp {i(θ1 Y (t) + θ2 Y (0))}] (4.10) − E [exp {iθ1 Y (t)}] E [exp {iθ2 Y (0)}] , θ1 , θ2 ∈ R, and I(t) : = I(θ1 , θ2 , t) : = − log E [exp {i(θ1 Y (t) + θ2 Y (0))}] + log E [exp {iθ1 Y (t)}] + log E [exp {iθ2 Y (0)}] , (4.11) θ1 , θ2 ∈ R. The following relationship between r(t) and I(t) is valid: ( ) r(θ1 , θ2 , t) = K(θ1 , θ2 , t) e−I(θ1 ,θ2 ,t) − 1 , where K(θ1 , θ2 , t) = E [exp {iθ1 Y (t)}] E [exp {iθ2 Y (0)}] = E [exp {iθ1 Y (0)}] E [exp {iθ2 Y (0)}] (4.12) := K(θ1 , θ2 ). Further, if I(t) → 0 as t → ∞, then r(t) ∼ −K(θ1 , θ2 )I(t) as t → ∞ which means r(t) and I(t) are asymptotically equivalent. If {Yt }t∈R is Gaussian, −I(1, −1, t) coincides with the covariance function and thus r(t) is a natural extension. The quantity r(t) was used in [2], where the authors studied the dependence structure of linear fractional stable motion. 59 Given a LTFSM (4.5), we define tempered fractional stable noise (TFSN) YH,α,λ (t) := XH,α,λ (t + 1) − XH,α,λ (t) for integers −∞ < t < ∞. (4.13) Remark 4.1.5. For two non-negative functions f (t) and g(t) on R, we will write f (t) ≍ g(t) f (t) if C1 ≤ g(t) ≤ C2 for all t sufficiently large, for some 0 < C1 < C2 < ∞. Theorem 4.1.6. Let 0 < α < 1, 0 < H < 1, λ > 0 and YH,α,λ (t) be the tempered fractional stable noise (4.13). Then r(θ1 , θ2 , t) ≍ e−λαt tHα−1 , as t → ∞. Proof. It follows easily from (4.5) that TFSN has the moving average representation ∫ +∞ [ ] H− 1 H− 1 YH,α,λ (t) = e−λ(t+1−x)+ (t + 1 − x)+ α − e−λ(t−x)+ (t − x)+ α Zα (dx). (4.14) −∞ 1 H− Define gt (x) = (t − x)+ α e−λ(t−x)+ for t ∈ R and compute I(θ1 , θ2 , t) and K(θ1 , θ2 , t) for TFSN {YH,α,λ } as follows: [ ] I(θ1 , θ2 , t) = − log E exp {i(θ1 YH,α,λ (t) + θ2 YH,α,λ (0))} [ ] [ ] + log E exp {iθ1 YH,α,λ (t)} + log E exp {iθ2 YH,α,λ (0)} ∫ +∞ α = θ1 [gt+1 (x) − gt (x)] + θ2 [g1 (x) − g0 (x)] dx − −∞ ∫ +∞ −∞ [ ]α dx − θ1 gt+1 (x) − gt (x) := I1 (θ1 , θ2 , t) + I2 (θ1 , θ2 , t), 60 ∫ +∞ −∞ [ ]α dx θ2 g1 (x) − g0 (x) (4.15) where ∫ 0 I1 (θ1 , θ2 , t) = − −∞ ∫ 0 −∞ θ1 [gt+1 (x) − gt (x)] + θ2 [g1 (x) − g0 (x)] [ ]α θ1 gt+1 (x) − gt (x) dx − ∫ 0 −∞ α dx [ ]α θ2 g1 (x) − g0 (x) dx and ∫ 1 α θ1 [gt+1 (x) − gt (x)] + θ2 g1 (x) dx 0 ∫ 1 ∫ 1 α [ ]α − θ1 gt+1 (x) − gt (x) dx − θ2 g1 (x) dx. I2 (θ1 , θ2 , t) = 0 0 Also, [ ] [ ] K(θ1 , θ2 ) = E eiθ1 Y (t) E eiθ2 Y (0) [ ] [ ] iθ Y (0) iθ Y (0) 1 2 =E e E e ∫ +∞ { } α α α = exp − (|θ1 | + |θ2 | ) g1 (x) − g0 (x) dx (4.16) −∞ by stationarity. Therefore, I(θ1 , θ2 , t) = K(θ1 , θ2 , t)(I1 (t) + I2 (t)) and to verify the asymptotic dependence of I(t) we just need to verify the asymptotic dependence of I1 (t) and I2 (t) as t → ∞. We first verify the asymptotic dependence of I1 (t). A change of variable in I1 (t) for t > 1 gives ∫ ∞ I1 (t) = 0 [ 1 1] θ1 e−λ(t+1+x) (t + 1 + x)H− α − e−λ(t+x) (t + x)H− α [ 1] α 1 + θ2 e−λ(1+x) (1 + x)H− α − e−λ(x) (x)H− α dx ∫ ∞ [ 1 1] α − θ1 e−λ(t+1+x) (t + 1 + x)H− α − e−λ(t+x) (t + x)H− α dx ∫0∞ [ −λ(1+x) 1 1] α H− −λ(x) H− α α − θ2 e (1 + x) dx. −e (x) 0 61 Let [ 1 1] α ft+1,t (x) := θ1 e−λ(t+1+x) (t + 1 + x)H− α − e−λ(t+x) (t + x)H− α . (4.17) For every t > 1 and x > 0 we get 1 eαλt t−α(H− α ) ft+1,t (x) = θ1 1 ( t + x )H− 1 α α −λ(1+x) ( t + 1 + x )H− α α e − e−λx t t α α → θ1 e−λαx e−λ − 1 as t → ∞ and α 1 sup eαλt t−α(H− α ) ft+1,t (x) ≤ θ1 (e−λ − 1) e−λαx , t>1 which belongs to L1 (0, ∞). Now we can use the Dominated Convergence Theorem to see that ∫ ∞ 0 α 1 ft+1,t (x) dx → θ1 (e−λ − 1) e−λαt tα(H− α ) = θ1 (e−λ − 1) α ∫ ∞ e−λαx dx 0 1 e−λαt tα(H− α ) λα (4.18) , as t → ∞. Now consider, gt,t+1,0,1 (x) := [ 1] 1 θ1 e−λ(t+1+x) (t + 1 + x)H− α − e−λ(t+x) (t + x)H− α [ 1 1] α + θ2 e−λ(1+x) (1 + x)H− α − e−λ(x) (x)H− α − θ2 α [ −λ(1+x) 1 1] α e (1 + x)H− α − e−λx (x)H− α . 62 (4.19) Then, 1 eλαt t−α(H− α ) g [ t,t+1,0,1 (x) = θ1 e−λ(1+x) ( 1 1 ) ) ( t + x H− α ] t + 1 + x H− α −λx −e t t [ ( 1 + x )H− 1 ( x )H− 1 ] α α α −λ(1+x) λt −λx λt e + θ2 e −e e t t where − ( 1 + x )H− 1 ( x )H− 1 ] α α α −λ(1+x) λt −λx λt θ2 e e −e e =: at + bt [ at = θ1 e−λ(1+x) and [ t ( α − bt t α 1 ) )H− 1 ] ( α t + 1 + x H− α t + x − e−λx t t [ ] ( 1 + x )H− 1 ( x )H− 1 α α −λ(1+x) λt −λx λt bt = θ2 e e −e e . t t It is obvious that at → Cx := θ1 e−λx (e−λ − 1) and bt → ∞ as t → ∞. Then, |at + bt |α − |bt |α → 0 as t → ∞ since 0 < α < 1. Therefore 1 eλαt t−α(H− α ) gt,t+1,0,1 → 0, as t → ∞. Moreover, for any 0 < α < 1, using the inequality |a|α − |b|α ≤ a − b α (see [61], Page 211), we get gt,t+1,0,1 ≤ ft+1,t , where gt,t+1,0,1 and gt,t+1,0,1 are defined in (4.17) and (4.19) respectively, if we let a = 63 θ1 (gt+1 − gt ) + θ2 (g1 − g0 ) and b = θ2 (g1 − g0 ). Consequently 1 1 sup eλαt t−α(H− α ) gt,t+1,0,1 ≤ sup eαλt t−α(H− α ) ft+1,t (x) t>1 t>1 ≤ θ1 (e−λ − 1) α −λαx e which also belongs to L1 (0, ∞). Applying the Dominated Convergence Theorem yields ∫ +∞ −∞ gt,t+1,0,1 (x)dx → 0 as t → ∞. (4.20) Therefore from (4.18) and (4.20) 1 |θ (e−λ − 1)|α e−λαt tα(H− α ) , I1 (t) ∼ − 1 λα (4.21) as t → ∞. Consider now, ∫ 1 α θ1 [gt+1 (x) − gt (x)] + θ2 g1 (x) dx 0 ∫ 1 ∫ 1 α α − θ1 [gt+1 (x) − gt (x)] dx − θ2 g1 (x) dx, I2 (t) = 0 0 Define, [ 1 1] ut (x) := θ1 e−λ(t+1−x) (t + 1 − x)H− α − e−λ(t−x) (t − x)H− α (4.22) and 1 v(x) := θ2 e−λ(1−x) (1 − x)H− α . We can rewrite ∫ 1 I2 (t) = 0 ξ(ut (x) + v(x)) − ξ(ut (x)) − ξ(v(x)) dx, 64 (4.23) where α ξ(g(x)) = g(x) . (4.24) Lemma 3.1 in [2] implies that ∫ 1 ξ(ut (x) + v(x)) − ξ(ut (x)) − ξ(v(x)) dx ∫ 1 α ut (x) dx, ≤ Pα I2 (t) = 0 (4.25) 0 where Pα = 2 + 4 tan( πα 2 )(see [2], Page 11). On the other hand ut (x) 1 1 1 − H)(t − x)H− α −1 e−λ(t−x) + λ(t − x)H− α e−λ(t−x) α [ ] 1 −1 1 1 H− α ≤ θ1 e−λ(t−1) ( − H) t − 1 + λ |t − 1|H− α α 1 [1 ] H− α −λ(t−1) ≤ θ1 e −H +λ t−1 , α θ1 ( (4.26) since 0 < x < 1. From (4.25) and (4.26) we get I2 (t) ≤ Pα ∫ 1 0 ut (x) α dx ]α [1 Hα−1 α −H +λ t−1 . = Pα θ1 e−λα(t−1) (4.27) α Hence from (4.21) and (4.27) we get I(t) ≍ e−λαt tHα−1 as t → ∞ and this completes the proof. Theorem 4.1.7. Let 1 < α < 2, α1 < H < 1, λ > 0 and let YH,α,λ (t) be the tempered 65 fractional stable noise (4.13). Then 1 I(t) ≍ e−λt t(H− α ) , as t → ∞. Proof. The proof is similar to that of Theorem 4.1.6. Let ft+1,t (x) be the function which is defined by (4.17). Then α 1 1 eλt t−(H− α ) ft+1,t (x) = θ1 eλt t−(H− α ) 1 α 1 e−λ(t+1+x) (t + 1 + x)H− α − e−λ(t+x) (t + x)H− α = at · bt , where α 1 at := θ1 eλt(α−1) t(H− α )(α−1) and ( ( x )H− α1 α 1 x )H− α1 − e−λx 1 + . bt := e−λ(1+x) 1 + + t t t Note that at → 0 (since 1 < α < 2) and bt → e−λ(1+x) − e−λx α as t → ∞. Now, let 1 h(t) = e−λt(α−1) t(α−1)(H− α ) . Observe that h(t) attains its maximum at t = λ1 (H − α1 ). Then α α 1 sup eλt t−(H− α ) ft+1,t (x) = e−λ − 1 θ1 e−λαx sup h(t) t>1 t>1 1) α −λαx −(H− 1 )(α−1) [ H − 1 ](α−1)(H− α α −λ α α e = e − 1 θ1 e , λ and so ft+1,t (x) is bounded by an L1 (0, ∞) function. Therefore the Dominated Convergence 66 Theorem implies that ∫ ∞ 0 ft+1,t (x) → 0 (4.28) 1 as t → ∞. Consider now, eλt t−(H− α ) gt,t+1,0,1 where gt,t+1,0,1 is given by (4.19). Then 1 eλt t−(H− α ) gt,t+1,0,1 = at + bt α − bt α where [ 1) 1) 1 ) −λ(1+x) ( t + 1 + x )(H− α 1 ) −λx ( t + x )(H− α −λt(1− −λt(1− α e α e at := θ1 e −e 1 1 tα tα and [ ] ] 1 λt −(H− α ) [ −λ(1+x) 1 1 ] α bt := θ2 e α t e (1 + x)(H− α ) − e−λx x(H− α ) . Observe that limt→∞ bt = ∞ and limt→∞ at = 0. Since |at + bt |α − |bt |α ∼ α|at ||bt |α−1 , as t → ∞, we get, 1 eλt t−(H− α ) gt,t+1,0,1 ∼ α θ1 1) 1) 1 ) −λx ( t + x )(H− α 1 ) −λ(1+x) ( t + 1 + x )(H− α −λt(1− −λt(1− α α −e e e e 1 1 tα tα α−1 λt(1− 1 ) −(H− 1 )(1− 1 ) −λ(1+x) 1 1 α−1 α t α α e θ2 e (1 + x)(H− α ) − e−λx x(H− α ) and consequently 1 eλt t−(H− α ) gt,t+1,0,1 → α θ1 e−λ(1+x) − e−λx θ2 α−1 −λ(1+x) 1 1 α−1 e (1 + x)H− α − e−λx xH− α . 67 Moreover, α α 1 sup eλt t−(H− α ) gt,t+1,0,1 = sup at + bt − bt t≥1 t≥1 ≤ sup at t≥1 α + α sup at bt α−1 (4.29) , t≥1 where we have used the following inequalities (Lemma 2 in [41]): |a − b|α ≤ aα + bα and |a + b|α − |b|α ≤ a α +α a b α−1 valid for a ≥ 0 and b ≥ 0 and α ∈ (1, 2). In order to find an upper bound for supt≥1 |at |α , write at α 1) 1) α α −λt(1− 1 ) −λ(1+x) ( t + 1 + x )(H− α 1 ) −λx ( t + x )(H− α −λt(1− α α e − e e e 1 1 tα tα 1) ( t + x )(H− 1 ) α α −λαx −λt(α−1) −λ ( t + 1 + x )(H− α α = θ1 e e e − 1 1 tα tα α −λαx −λ α 1 1 ≤ θ1 e e (1 + 1 + x)H− α − (1 + x)H− α = θ1 ] α −λαx [ −λα Hα−1 Hα−1 ≤ θ1 e e (2 + x) + (1 + x) α ≤ 2 θ1 e−λαx (2 + x)Hα−1 . (4.30) On the other hand, α at bt α−1 = α θ1 θ2 α−1 ( t + x )(H− 1 ) ( t + 1 + x )(H− 1 ) α α −λx −λ(1+x) −e × K(x) e t t :=S(t) where 1 1 K(x) = e−λ(1+x) (1 + x)(H− α ) − e−λx (x)(H− α ) 68 α−1 . (4.31) Note that S(t) is a decreasing function and hence sup α at bt α−1 t≥1 = α θ1 θ2 α−1 (4.32) 1 1 e−λ(1+x) (2 + x)(H− α ) − e−λx (1 + x)(H− α ) × K(x) where K(x) is given by (4.31). From (4.29), (4.30) and (4.32) α α−1 1 sup eλt t−(H− α ) gt,t+1,0,1 ≤ 2 θ1 e−λαx (2 + x)Hα−1 + α θ1 θ2 t≥1 (4.33) 1 1 e−λ(1+x) (2 + x)(H− α ) − e−λx (1 + x)(H− α ) × K(x) which belongs to L1 (0, ∞), since Hα > 1. Then, the Dominated Convergence Theorem implies that ∫ ∞ 0 α−1 1 gt,t+1,0,1 (x) dx → αθ1 θ2 e−λt t(H− α ) ∫ ∞ 1 1 α−1 e−λ(1+x) − e−λx e−λ(1+x) (1 + x)H− α − e−λx xH− α dx 0 1 = C2 (α, λ, θ1 , θ2 )e−λt t(H− α ) , (4.34) as t → ∞, where α−1 C2 (α, λ, θ1 , θ2 ) = αθ1 θ2 (4.35) ∫ ∞ 1 1 α−1 −λ(1+x) H− −λ(1+x) −λx −λx H− α α e (1 + x) e −e dx −e x 0 is a constant that is independent of t. Therefore from (4.28) and (4.34) 1 I1 (t) ∼ C2 (α, λ, θ1 , θ2 )e−λt t(H− α ) , 69 (4.36) as t → ∞. Finally, recall that ∫ 1 α θ1 [gt+1 (x) − gt (x)] + θ2 g1 (x) dx 0 ∫ 1 ∫ 1 α α − θ1 [gt+1 (x) − gt (x)] dx − θ2 g1 (x) dx, I2 (t) = 0 0 and that ut (x) and v(x) are given by (4.22) and (4.23) respectively. Then ∫ 1 I2 (t) = 0 ξ(ut (x) + v(x)) − ξ(ut (x)) − ξ(v(x)) dx where ξ(g(x)) is given by (4.24). Lemma 3.1 in [2] implies that, using argument similar to (4.26), we have ∫ 1 ξ(ut (x) + v(x)) − ξ(ut (x)) − ξ(v(x)) dx ∫ 1 ∫ 1 α−1 α ≤ Rα ut (x) v(x) dx + Sα ut (x) dx I2 (t) = 0 0 ≤ Rα θ1 ∫ 1[ 0 1 α−1 H− 1 α −λ(t−1) e v(x) dx H − +λ t−1 ] α 0 [ ]α α Hα−1 −λα(t−1) 1 + Sα θ1 H − + λ t − 1 e α 1 ] [ H− α 1 = Rα θ1 H − + λ t − 1 e−λ(t−1) α ∫ 1 1 α−1 · θ2 e−λ(1−x) (1 − x)H− α dx 0 + Sα θ1 α[ ]α Hα−1 −λα(t−1) 1 H − +λ t−1 e α 1 H− α := C3 (α, λ, θ1 ) t − 1 e−λ(t−1) + Sα θ1 α[ H− ]α Hα−1 −λα(t−1) 1 e , +λ t−1 α 70 (4.37) where [ ] 1 C3 (α, λ, θ1 ) = Rα θ1 |H − | + λ α ∫ 1 1 α−1 · θ2 e−λ(1−x) (1 − x)H− α dx 0 ) ( πα is a constant. Recall that Rα = α 1 + tan( πα 2 ) , and Sα = (α + 1) + (α + 3) tan( 2 ) (see [2], Page 11) are also constants. Note that the upper bound which is obtained by (4.37) is of the same order as the upper bound for I1 (t), given by (4.34). Hence 1 I(t) ≍ e−λt t(H− α ) and this completes the proof. Definition 4.1.8. A symmetric α-stable stationary process {Yt } has long memory if r(θ1 , θ2 , t) defined in (4.10) satisfies ∞ ∑ r(θ1 , θ2 , n) = ∞ (4.38) n=0 Lemma 4.1.9. The LTFSM process does not have long memory property in the sense of (4.38). Proof. From Theorems 4.1.6 and 4.1.7 we have ∞ ∑ r(θ1 , θ2 , n) < ∞ n=0 which proves the statement. Remark 4.1.10. Theorem 4.1.7 gives the fact that when the tempering parameter λ is sufficiently small, TFSN exhibits semi-long range dependence, with the asymptotic rate I(t) 71 1 that falls off like tH− α for moderate values of t > 1, and eventually that rate falls off like 1 e−λt tH− α for t sufficiently large. 4.2 Tempered fractional harmonizable stable motion Let X = X1 + iX2 be a complex-valued random variable. We say X is isotropic SαS if the vector (X1 , X2 ) is SαS and for any θ = θ1 + iθ2 E [exp {i(θ1 X1 + θ2 X2 )}] = exp {−c|θ|α } , for some constant c > 0. A complex-valued stochastic process {X(t)} is called isotropic SαS if all complex linear combinations ∑n j=1 θj X(tj ) are complex-valued isotropic SαS random variables. Let (R, B(R), dx), where B(R) is the σ-algebra of Borel subsets of R. We say the process {Zα (B), B ∈ B(R)} is a complex-valued isotropic SαS random measure with Lebesgue control measure dx if [ { }] E exp iRe(θZα (B)) = exp {−|B||θ|α } , for any θ ∈ C where |B| denotes the Lebesgue measure of B (B ∈ B(R)). Definition 4.2.1. Let {Zα (B), B ∈ B(R)} be a complex-valued isotropic SαS random measure with Lebesgue control measure dx. Then the stochastic integral ∫ +∞ I(f ) := Re −∞ 72 f (k)Zα (dk), where f ∈ Lα (R) is the complex-valued SαS random variable such that [ { E exp iRe ∫ +∞ −∞ }] { ∫ +∞ } α f (k)Zα (dk) = exp − f (k) (dk) −∞ and I(f ) α α ∫ +∞ = Re ( := −∞ f (k)Zα (dk) α [ ]) − log E exp{iRe I(f )} ∫ +∞ = α −∞ f (k) α (4.39) dk, for any 0 < α < 2. Definition 4.2.2. The real harmonizable tempered fractional stable motion (HTFSM) is the process ∫ +∞ XH,α,λ (t) = Re e−ikt − 1 1 −∞ (λ − ik)H+ α Zα (dk) (4.40) where 0 < α < 2, H > 0, λ > 0 and Z˜α is a complex isotropic SαS random measure. Remark 4.2.3. The stochastic integral in (4.40) is well defined, since ∫ +∞ −∞ α e−ikt − 1 1 (λ − ik)H+ α dk < ∞, (4.41) for any H > 0 and 0 < α < 2 (as |k| → ∞, the integrand behaves like |k|−Hα−1 , which is integrable for any 0 < α < 2 and H > 0; as |k| → 0, the integrand tends to zero). Definition 4.2.4. Given a HTFSM (4.40), we define tempered fractional harmonizable sta- 73 ble noise (TFHSN) YH,α,λ (t) := XH,α,λ (t + 1) − XH,α,λ (t) for integers −∞ < t < ∞. (4.42) It follows easily from (4.40) that TFHSN has the harmonizable representation ∫ +∞ YH,α,λ (t) = Re −∞ eikt Ψ(dk), (4.43) Zα (dk) (4.44) where Ψ(dk) = eik − 1 1 (λ + ik)H+ α is a complex symmetric α-stable (SαS) random measure with the control measure m(dk) = |eik − 1|α dk, |λ + ik|Hα+1 (4.45) for any 0 < α < 2, λ > 0 and H > 0. Theorem 4.2.5. The tempered fractional stable motion (LTFSM) defined in (4.5) and tempered fractional harmonizable stable motion (HTFSM) defined in (4.40) are different processes. Proof. To prove that the processes {XH,α,λ (t)} and {XH,α,λ (t)} are different it is enough to show that lim rY t→∞ (θ1 , θ2 , t) = 0 (4.46) (θ1 , θ2 , t) > 0, (4.47) H,α,λ and lim r t→∞ YH,α,λ 74 where {YH,α,λ (t)} and {YH,α,λ (t)} are the increments of {XH,α,λ (t)} and {XH,α,λ (t)} respectively. Lemma 6.1 in [33] shows that if ∫ +∞ Y (t) = −∞ f (t − x)Zα (dx), for f ∈ Lα (R), {Zα } is the SαS random measure on R, then lim rY (θ1 , θ2 , t) = 0. t→∞ 1 1 H− H− Along the same lines, we now define f (x) := (x+1)+ α e−λ(x+1)+ −(x)+ α e−λ(x)+ . Then YH,α,λ (t) = XH,α,λ (t + 1) − XH,α,λ (t) ∫ +∞ [ ] H− 1 H− 1 = (t − x + 1)+ α e−λ(t−x+1)+ − (t − x)+ α e−λ(t−x)+ Zα (dx) −∞ ∫ +∞ = −∞ f (t − x) Zα (dx) and hence by applying Lemma 6.1 [33], as described above, lim rY t→∞ H,α,λ (θ1 , θ2 , t) = 0 which is Property (4.46). Theorem 3.1 in [37] states that if ∫ +∞ Y (t) = Re −∞ eikt Ψ(dk) for −∞ < t < ∞ 75 is a stationary real harmonizable SαS process, then ( ) ∫ T 1 1 lim inf r (θ1 , θ2 , t) dt ≥ K(θ1 , θ2 )c0 m{0}F0 + m(R − {0})F1 > 0, 2π T →∞ 2T −T Y where m is the control measure of the isotropic complex-valued random measure Ψ, F0 ∈ R and F1 > 0 are some constants depends on α, m, θ1 and θ1 . Along the same lines, we now define YH,α,λ (t) = XH,α,λ (t + 1) − XH,α,λ (t) ∫ +∞ −ik(t+1) e − e−ikt = Re Zα (dk) 1 −∞ (λ − ik)H+ α ∫ +∞ eikt Ψ(dk), = Re −∞ where Ψ(dk) and m(dk) are given by (4.44) and (4.45) respectively. Hence by applying Theorem 3.1 in [37], as explained above, we get ( ) ∫ T 1 1 r (θ1 , θ2 , t) dt ≥ K(θ1 , θ2 )c0 m{0}F0 + lim inf m(R − {0})F1 > 0 2π T →∞ 2T −T YH,α,λ and consequently lim r t→∞ YH,α,λ (θ1 , θ2 , t) > 0, which is Property (4.2) and this completes the proof of the theorem. 4.3 Sample path properties In this section, we present some results about the sample path properties of LTFSM and HTFSM. The path behavior of the linear fractional stable motion (LFSM) process XH,α 76 50 20 X(t) 0 0 −100 −50 −40 −80 X(t) 0 100 200 300 400 500 0 t 100 200 300 400 500 t Figure 4.1: Left panel: Sample paths of LTFSM (thick black line) with λ = 0.03 and H = 0.3, and LFSM (thin black line) with H = 0.3. Both graphs use the same noise realization Zα (t). The right panel shows the same plots for λ = 0.001, H = 0.7 and α = 1.5. 1 H− (see Definition 7.4.1 [61]) depends on the structure of the kernel gα,t (x) := (t − x)+ α − 1 H− (−x)+ α , t, x ∈ R (see [62]). When H − α1 < 0, the function gα,t (x), x ∈ R, has singu- larities at x = 0 and x = t. By the same argument, The paths behavior of the LTFSM 1 H− process XH,α,λ depends on the structure of the kernel gα,λ,t (x) := (t − x)+ α e−λ(t−x)+ − 1 H− (−x)+ α e−λ(t−x)+ . In fact, The function gα,λ,t (x) has singularities at x = 0 and x = t too. These singularities magnify the stable noise processes Zα (dx) and cause large spikes in the paths of the LFSM and LTFSM processes. The left panel in Figure 4.1 compares a typical sample path of both processes, simulated using the same noise realization Zα (t), in the case H − α1 < 0. In the case H − α1 > 0, (since 0 < H < 1 it follows that α > 1) the paths of the fractional stable motion can be made continuous with probability one (see Chapter 10 in [61]), since its kernel is bounded and positive for all t > 0. Similarly, the kernel of LTFSM, gα,λ,t (x), is bounded and positive for all t > 0 and hence the paths of the process XH,α,λ can be made continuous with probability one when H − α1 > 0. The right panel in Figure 4.1 shows the corresponding sample paths in the case α = 1.5. These simulations use a discretized version of the moving average representation of LTFSM (4.5). Theorem 4.3.1. Let 0 < α < 2, 0 < H < α1 and X = {XH,α,λ (t)}t∈R be the LTFSM 77 ∗ process defined in (4.5). Then, for any separable version X ∗ = {XH,α,λ (t), t ∈ (a, b)} of the process X, we have that ( P {ω : sup t∈(a,b) ) ∗ (t, ω) = ∞} = 1, XH,α,λ That is, every version of the process X = {XH,α,λ (t)}t∈R has unbounded paths. Proof. We apply Theorem 10.2.3 in [61]. Indeed, consider the countable set T ∗ := Q ∩ [a, b], where Q denotes the set of rational numbers. Since T ∗ is dense in [a, b], there exists a sequence {tn }n∈N ∈ T ∗ such that tn → x as n → ∞. Therefore f ∗ (T ∗ ; x) := sup gα,λ,t (x) t∈T ∗ ≥ sup gα,λ,tn (x) =: fn∗ (T ∗ ; x) = ∞, tn ∈T ∗ ∫ as n → ∞ and hence ab f ∗ (T ∗ ; x) dx = ∞, and this contradicts Condition (10.2.14) of Theorem 10.2.3 in [61]. Therefore, the stochastic process {XH,α,λ } does not have a version with bounded paths on the interval (a, b), and this completes the proof. Lemma 4.3.2. Let 1 < α < 2, α1 < H < 1 and λ > 0. Then there exist a positive constant C1 such that the LTFSM (4.5) satisfies XH,α,λ (t) − XH,α,λ (s) locally uniformly in s, t ∈ [0, 1]. 78 α α ≥ C1 t − s Hα , Proof. We write XH,α,λ (t) − XH,α,λ (s) α ∫ t 1 |t − u|α(H− α ) e−λα|t−u| du α s ∫ t −λα|t−s| ≥e |t − u|Hα−1 du ≥ s = e−λα|t−s| Hα |t − s|Hα = C1 |t − s|Hα and this gives the lower bound. Lemma 4.3.3. Let 1 < α < 2, α1 < H < 1 and λ > 0. Then there exist positive constants C1 and C2 such that the HTFSM (4.40) satisfies C1 t − s Hα α ≤ XH,α,λ (t) − XH,α,λ (s) ≤ C2 t − s α Hα , (4.48) locally uniformly in s, t ∈ [0, 1]. Proof. To get the upper bound, write ∫ +∞ −ikt |e − e−iks |α dk XH,α,λ (t) − XH,α,λ (s) = α |λ − ik|Hα+1 −∞ ∫ +∞ ≤C (1 ∧ |t − s|α |k|α ) |λ − ik|−Hα−1 dk α −∞ ∫ [ α = C |t − s| |k|< 1 |t−s| ∫ + |k|> 1 |t−s| |k|α ||λ − ik|−Hα−1 dk |λ − ik|−Hα−1 dk ] [ ≤ C |t − s|α I1 + I2 79 ] (4.49) where ∫ I1 := |k|< 1 |t−s| and α k λ − ik ∫ I2 := −Hα−1 λ − ik |k|> 1 |t−s| −Hα−1 dk dk and C is a constant. Observe that ∫ I1 = |k|< 1 |t−s| k ∫ ≤ |k|< 1 |t−s| ≤ t−s k Hα−α α α · −Hα−1 2 λ2 + k 2 k dk ∫ −Hα−1 dk = |k|< 1 |t−s| k −Hα−1+α dk (4.50) 2 α(1 − H) and ∫ I2 = λ2 + k 2 |k|> 1 |t−s| ∫ ≤ k2 |k|> 1 |t−s| ≤ t−s Hα · −Hα−1 2 dk ∫ −Hα−1 2 dk = |k|> 1 |t−s| k −Hα−1 dk (4.51) 2 . Hα Finally, from (4.49), (4.50) and (4.51) we get XH,α,λ (t) − XH,α,λ (s) α α ≤C [ [ α t − s I1 + I2 ] 2 2 ≤C + α(1 − H) Hα = C2 t − s ] t−s Hα Hα which gives the upper bound in (4.48). In order to get the lower bound, we use the fact that 80 there exist positive constants c1 , c2 such that |e−iy − 1| > c1 |y| for |y| < c2 . Therefore XH,α,λ (t) − XH,α,λ (s) α α ∫ +∞ = −∞ ∫ +∞ = −∞ e−ikt − e−iks α e−ik(t−s) − 1 α ∫ ≥ c1 |k|< c2 |t−s| k α λ − ik α t−s ∫ = c1 |t − s|α λ − ik c2 |t−s| −(Hα+1) λ − ik α |k|< −(Hα+1) dk dk −(Hα+1) k (λ2 + k 2 ) dk −(Hα+1) 2 dk. We now use the fact that ( λ2 + k 2 ) −(Hα+1) 2 ≥ ( 1 + c22 ) −(Hα+1) 2 t−s Hα+1 , c 1 and |k| < 2 to continue the rest of the proof as follows: for λ < |t−s| |t−s| c1 t − s α ∫ ) α( 2 k λ + k2 −(Hα+1) 2 dk c2 |t−s| ∫ c2 ) −(Hα+1) ( α Hα+1 |t−s| α 2 2 k dk ≥ 2c1 1 + c2 t−s t−s 0 Hα+α+1 −α−1 Hα = C1 t − s t−s = C1 t − s |k|< and this gives the lower bound. 4.4 Local Times and Local nondeterminism In this section, we prove the existence of the local time for LTFSM and HTFSM. We also show that LTFSM and HTFSM are locally nondeterministic on every compact interval. We first recall the definition of the local time (see [11] for more details). Suppose X = {X(t), t ≥ 0} 81 is a real-valued separable random process with Borel sample functions. For any Borel set B ⊂ R+ µB (A) = η({s ∈ B, X(s) ∈ A}) is called the occupation measure of X on B, where η is the Lebesgue measure on R+ . If µB is absolutely continuous with respect to the Lebesgue measure on R, then we say that X has a local time on B and define its local time, L(B, .), to be the Radon-Nikodym derivative of µB with respect to Lebesgue measure. We write L(t, x) instead of L([0, t], x). Proposition 4.4.1. The LTFSM defined in (4.5) with 1 < α < 2 and α1 < H < 1 has a square integrable local time L(t, x). Proof. According to Theorem 3.1 in [11], a stochastic process {X(t), t ∈ [0, T ]} has a local time L(t, x) which is continuous in t for a.e. x ∈ R and square integrable with respect to x if {X(t), t ∈ [0, T ]} satisfies • Condition (H): There exist positive numbers (ρ0 , H) ∈ (0, ∞) × (0, 1) and a positive function ψ ∈ L1 (R) such that for all κ ∈ R, t, s ∈ [0, T ], 0 < |t − s| < ρ0 we have ( X(t) − X(s) ) E exp iκ ≤ ψ(κ). |t − s|H (4.52) We prove that the LTFSM {XH,α,λ (t)} satisfies (H). Apply (4.7) and Lemma 4.3.2 to get α) ) ( ( X H,α,λ (t) − XH,α,λ (s) α ∥XH,α,λ (t) − XH,α,λ (s)∥α = exp − |κ| E exp iκ |t − s|H |t − s|αH ( ) ≤ exp − |κ|α C := ψ(κ) where the function ψ(κ) belongs to L1 (R, dk) which means the HTFSM satisfies H and this 82 completes the proof. Proposition 4.4.2. The HTFSM defined in (4.40) with 1 < α < 2 and α1 < H < 1 has a square integrable local time L(t, x). Proof. We prove that the HTFSM {XH,α,λ (t)} satisfies (H). Apply (4.39) and Lemma 4.3.3 to obtain α) ( X ) ( H,α,λ (t) − XH,α,λ (s) α ∥XH,α,λ (t) − XH,α,λ (s)∥α E exp iκ = exp − |κ| |t − s|H |t − s|αH ( ) ≤ exp − |κ|α C := ψ(κ), where the function ψ(κ) belongs to L1 (R, dk) which means the HTFSM satisfies H and this completes the proof. We next show that HTFSM is locally nondeterministic on every compact interval [ϵ, T ], for any 0 < ϵ < T < ∞. Recall that a stochastic process {X(t)}t∈T is locally nondeterministic (LND) if 1. ∥X(t)∥α > 0 for all t ∈ T 2. ∥X(t) − X(s)∥α > 0 for all t, s ∈ T sufficiently close; and 3. for any m ≥ 2, lim inf ϵ↓0 ∥X(tm ) − span{X(t1 ), . . . , X(tm−1 )}∥α > 0, ∥X(tm ) − X(tm−1 )∥α where the lim inf is taken over distinct, ordered t1 < t2 < . . . < tm ∈ T with |t1 − tm | < ϵ, [ ]1/α T ⊂ R and ∥X∥α = − log(E exp(iX)) for 0 < α ≤ 2 (see [51, 52, 67] for more details). Next, we show that HTFSM also is LND. 83 Proposition 4.4.3. The HTFSM (4.40) with 1 < α < 2 and α1 < H < 1 is LND on every interval [ϵ, κ] for ϵ < κ < ∞. Proof. Theorem 3.3 in [19] shows that a harmonizable multifractional stable motion is LND [ϵ, κ] for ϵ < κ < ∞. Our proof is a modification of that Theorem. We need to verify conditions (1), (2) and (3) as described above. The first and second condition follows from Lemma 4.3.3 α XH,α,λ (t) − XH,α,λ (s) α ≥ C1 t − s Hα , where C1 is a positive constant. To prove the third condition, observe first that the inverse Fourier transform of fH,α,λ (t, k) := e−ikt − 1 (4.53) 1 (λ − ik)H+ α on Lα (R) which is F −1 fH,α,λ (t, k) = α−1 ] Γ(H + α1 ) [ −λ(t−x) H− α−1 α − e−λ(−x)+ (−x)H− α + (t − x) √ e + + 2π (4.54) by (2.11). In order to verify the third condition, we shall establish a lower bound for XH,α,λ (tm ) − m−1 ∑ j=1 uj XH,α,λ (tj ) = fH,α,λ (tm , k) − α m−1 ∑ j=1 uj fH,α,λ (tj , k) Lα (R) α . By applying the Hausdorff-Young where fH,α,λ (t, k) is defined in (4.53). Let β = α−1 84 inequality (see Theorem 5.7 in [39]): fH,α,λ (tm , k) − m−1 ∑ uj fH,α,λ (tj , k) Lα (R) j=1 ≥ C F −1 fH,α,λ (tm , k) − uj F −1 fH,α,λ (tj , k) j=1 (∫ =C m−1 ∑ tm−1 −∞ ∫ tm + tm−1 F −1 fH,α,λ (tm , k) − m−1 ∑ Lβ (R) (4.55) uj F −1 fH,α,λ (tj , k) β j=1 F −1 fH,α,λ (tm , k) β )1 β dk . From (4.54) we have F −1 fH,α,λ (tm , k) = α−1 ] Γ(H + α1 ) [ −λ(tm −x) H− α−1 α − e−λ(−x)+ (−x)H− α + (tm − x) √ e + + 2π α−1 H− and the second term, e−λ(−x)+ (−x)+ α , vanishes on the interval [tm−1 , tm ]. Hence we can continue (4.55) as the following: [∫ ≥C tm tm−1 ]1 1 β β(H− ) −λβ(tm −x) β e dx (tm − x) (4.56) H ≥ C e−λ(tm −tm−1 ) tm − tm−1 ≥ C e−λ(κ−ϵ) XH,α,λ (tm ) − XH,α,λ (tm−1 ) α for tm and tm−1 close enough (and C is a constant). In the last line in (4.56), we used the fact that |tm − tm−1 | < κ − ϵ and we also applied Lemma 4.3.3 to get the last inequality. 85 Therefore XH,α,λ (tm ) − span{XH,α,λ , . . . , XH,α,λ (tm−1 )} α = XH,α,λ (tm ) − m−1 ∑ uj XH,α,λ (tj ) j=1 α ≥ C XH,α,λ (tm ) − XH,α,λ (tm−1 ) α and consequently lim inf ϵ↓0 XH,α,λ (tm ) − span{XH,α,λ , . . . , XH,α,λ (tm−1 )} XH,α,λ (tm ) − XH,α,λ (tm−1 ) α > C, α where C is a positive constant and this verifies Condition (3) of the LND property and this completes the proof. 86 BIBLIOGRAPHY 87 BIBLIOGRAPHY [1] Adler, R., Feldman, R., Taqqu, M.: A Practical Guide to Heavy Tails: Statistical Techniques and Applications. Springer (1998). [2] Astrauskas, A., Levy, J., B., Taqqu, M., S., The asymptotic dependence structure of the linear fractional Levy motion, Lietuvos Matematikos Rinkinys (Lithuanian Mathematical Journal), 31 (1) , 1–28(1991) [3] Baeumer, B., Meerschaert, M.M.: Stochastic solutions for fractional Cauchy problems. Fractional Calculus and Applied Analysis, 4, 481–500 (2001). [4] Baeumer, B., Meerschaert, M.M.: Tempered stable L´evy motion and transient superdiffusion. Journal of Computational and Applied Mathematics, 233, 2438–2448 (2010). [5] Banerjee, S., Gelfand, A.E.: On Smoothness Properties of Spatial Processes. Journal of Multivariate Analysis, 84, 85–100 (2003). [6] Barton, R.J., Poor, V.H.: Signal Detection in Fractional Gaussian Noise. IEEE Transactions on Information Theory, 34(5), 943–955 (1998). [7] Beran, J.: Statistics for long-memory processes. Chapman & Hall/CRC (1994). [8] Berman, S.M.: Local nondeterminism and local times of Gaussian processes. Indiana Univ. Math. J. 23, 69-94 (1973). [9] Biagini, F., Hu, Y., Øksendal, B., Zhang T.: Stochastic Calculus for Fractional Brownian Motion and Applications. Springer (2010). [10] Bonami, A., Estrade, A.: Anisotropic analysis of some Gaussian models. J. Fourier Anal. Appl. 9, 215–236 (2003). [11] Boufoussi, B., Dozzi, M. E., Guerbaz, R., Path properties of a class of locally asymptotically self similar processes. Electron. J. Probab. 13, 898-921 (2008). ´ del-Castillo-Negrete, D.: Fluid limit of the continuous-time random walk [12] Cartea, A., with general L´evy jump distribution functions. Physical Review E 76, 041105 (2007). 88 [13] Chakrabarty, A., Meerschaert, M.M.: Tempered stable laws as random walk limits. Statistics and Probability Letters 81(8), 989–997 (2011). [14] Cheridito, P.: Gaussian moving averages, semimartingales and option pricing. Stochastic Processes and their Applications, 109(1), 47–68 (2004). [15] Dalang, R.C., Khoshnevisan, D., Rassoul-Agha, F.: A Minicourse on Stochastic Partial Differential Equations. Lecture Notes in Mathematics 1962, Springer (2009). [16] Davenport, A.G.: The spectrum of horizontal gustiness near the ground in high winds. Q. J. Royal Meteor. Soc. 87, 194–211 (1961). [17] Demengel, F., Demengel, G: Functional Spaces for the Theory of Elliptic Partial Differential Equations. Universitext 8, Springer (2012). [18] Doukhan, P., Oppenheim, G., Taqqu, M.S.: Theory and Applications of Long-Range Dependence. Springer (2003). [19] Dozzi, M., Shevchenko, G., Real harmonizable multifractional stable process and its local properties Stochastic Processes and their Applications, 121 (7), 1509–1523 (2011). [20] Dozzi, M., Soltani, A. R., Stochastic Processes and their Applications, 68 (2), 195–207 (1997). Marco Dozzi, [21] E.Barndorff-Nielsen, O.: Processes of normal inverse Gaussian type. Finance Stoch. 2, 41–68 (1998). [22] Embrechts, P., Maejima, M.:Selfsimilar Processes. Princeton University Press (2002). [23] Gneiting, T., Kleiber, W., Schlather, M.: Mat´ern cross-covariance functions for multivariate random fields. Journal of the American Statistical Association. 105, 1167–1177 (2010). [24] Gradshteyn, I.S., Ryzhik, I.M.: Table of Integrals and Products. Sixth edition, Academic Press (2000). [25] Grenander, G.: Abstract Inference. Wiley (1981). [26] Handcock, M.S., Stein, M.L.: A Bayesian analysis of kriging. Technometrics 35, 403–410 (1993). 89 [27] Huang, S.T., Cambanis, S.: Stochastic and multiple Wiener integrals for Gaussian processes. The Annals of Probability 6, 585–614 (1978). [28] Jang, J.-J., Jyh-Shinn, G.: Analysis of maximum wind force for offshore structure design. J. Marine Sci. Tech. 7(1), 43–51(1999). [29] Jolis, M.: The Wiener integral with respect to second order processes with stationary increments. Journal of Mathematical Analysis and Applications 336, 607–620 (2010). [30] Kallenberg, O.: Foundations of Modern Probability. Second edition, Springer, New York (2002). [31] Kierat, W., Sztaba, U.: Distributions, Integral Transforms and Applications. Taylor and Francis, CRC Press (2003). [32] Knight, F.: Foundations of the prediction process. Oxford, Clarendon Press (1992). [33] Kokoszka, P. S., Taqqu, M., S., New classes of self-similar symmetric stable random fields J. Theoret. Probab., 7, 527-549 (1994). [34] Kolmogorov, A.N.: Wiener spiral and some other interesting curves in Hilbert space. Dokl. Akad. Nauk SSSR 26, 115–118 (1940). [35] Kˆono, N., Maejima, M., Self-similar stable processes with stationary increments, in: Stable processes and related topics, Sel. Pap. Worksop, Ithaca/NY (USA) 1990, in: Prog. Probab., 25, 275–295 (1991). [36] Kuo, H.H.: White noise distribution theory. CRC Press, Boca Raton, Florida (1996). [37] Levy, J., and Taqqu, M. S., A characterization of the asymptotic behavior of stationary stable processes. In Samorodnitsky, G., Cambanis, S., and Taqqu, M. S.,(eds.),Stable Processes and Related Topics, Boston, Birkhuser,Progress in Probability 25, 181-198 Levy,(1991). [38] Li, Y., Kareem, A.: ARMA systems in wind engineering. Probabilistic Engineering Mechanics 5, 49–59 (1990). [39] Lieb, E., H., Loss, M., Analysis, Graduate Studies in Mathematics (2nd ed.), vol. 14, American Mathematical Society (AMS), Providence, RI (2001) xxii, 346 p. 90 [40] Lim, S.C., Teo, L.P.: Weyl and Riemann-Liouville multifractional Ornstein–Uhlenbeck processes. Journal of Physics A 40, 6035–6060 (2007). [41] Magdziarz, M., The dependence structure of the fractional Ornstein-Uhlenbeck process, Prob. and Math. Stat. 25, 97–104(2005). [42] Magdziarz, M., Weron, A., Fractional langevin equation with α-stable noise. A link to fractional ARIMA time series, Stud. Math., 181, 47(2007). [43] Mandelbrot, B.:The Fractal Geometry of Nature, Freeman, San Francisco (1982). [44] Mandelbrot, B., Van Ness, J.: Fractional Brownian motion, fractional noises and applications. SIAM Review 10, 422–437 (1968). [45] Mandrekar, V., Meerschaert, M.M., Sabzikar, F.: RKHS representation for tempered fractional Brownian motion. Preprint (2013). [46] Meerschaert, M.M., Sabzikar, F.: Tempered fractional Brownian motion. Statistics and Probability Letters 83(10), 2269–2275 (2013). [47] Meerschaert, M.M., Scheffler, H.-P. Limit Distributions for Sums of Independent Random Vectors: Heavy Tails in Theory and Practice. Wiley Interscience, New York (2001). [48] Meerschaert, M.M., Sikorskii, A.: Stochastic Model For Fractional Calculus. de Gruyter, Berlin/Boston (2012). [49] Meerschaert, M.M., Zhang, Y., Baeumer, B.: Tempered anomalous diffusion in heterogeneous systems. Geophysical Research Letters 35, L17403 (2008). [50] Nezza, E., Di., Palatucci, G., and Valdinoci, E.: Hitchhiker’s guide to the fractional Sobolev spaces. Bull. Sci. Math. 136, 521–573 (2012). [51] Nolan, J., P.: Path properties of index-β stable fields. Annals of probability, Vol 16, No. 4, pp. 1596–1607, (1988). [52] Nolan, J., P.: Local nondeterminism and local times for stable processes. Probab. Theory Related Fields 82 , No. 3, pp. 387–410. MR1001520 (91b:60060),(1989). [53] Norton, D.J.: Mobile Offshore Platform Wind Loads. Proc. 13th Offshore Techn. Conf., OTC 4123, 4, 77–88 (1981). 91 [54] Oldham, K.B., Spanier, J.: The Fractional Calculus. Academic Press (1974). [55] P´erez Beaupuits, J.P., Ot´arola, A., Rantakyr¨o, F.T., Rivera, R.C., Radford, S.J.E., Nyman, L.-˚ A.: Analysis of wind data gathered at Chajnantor. ALMA Memo 497 (2004). [56] Pipiras, V., Taqqu, M.: Integration questions related to fractional Brownian motion, Probability Theory and Related Fields 118, 251–291 (2000). [57] Pitman, E.J.G.: On the behaviour of the characteristic function of a probability distribution in the neighbourhood of the origin. J. Austral. Math. Soc 8, 422–443(1968). [58] Rangarajan, G., Ding, M.: Processes with Long-Range Correlations: Theory and Applications. Springer (2003). [59] Rosi´ nski, J.: Tempering stable processes. Stochastic Processes and their Applications, 117, 677–707 (2007). [60] Samko, S.G, Kilbas, A.A., Marichev, O.I.: Fractional Integrals and Derivatives. Gordon and Breach (1993). [61] Samorodnitsky, G., Taqqu, M.: Stable Non-Gaussian Random Processes: Stochastic Models with Infinite Variance. Chapman & Hall (1994). [62] Stoev, S., Taqqu., M., S.,Simulation methods for linear fractional stable motion and FARIMA using the fast Fourier transform, Fractals, (2004) [63] Tudor, C.: Inner product spaces of integrands associated to subfractional Brownian motion. Statistics and Probability Letters 78, 2201-2209 (2008). [64] Tudor, C.: Analysis of Variations for Self-similar Processes: A Stochastic Calculus Approach. Springer (2013). [65] Weinert, H., L.: Reproducing Kernel Hilbert Spaces: Applications in Statistical Signal Processing. Hutchinson Ross (1982). [66] Wheeden, R.L., Zygmund, A.: Measure and Integral. Marcel Dekker, New York (1977). [67] Xiao, Y., Properties of local nondeterminism of Gaussian and stable random elds and their applications. Ann. Fac. Sci. Toulouse Math. XV, pp. 157-193 (2006). 92 [68] Yosida, K.: Functional Analysis. Sixth edition, Springer (1980). [69] Zhang, Y., Benson, D.A., Meerschaert, M.M., Scheffler, H.-P.: On using random walks to solve the space-fractional advection-dispersion equations. Journal of Statistical Physics 123, 89–110 (2006). [70] Zhang, Y, Meerschaert, M.M.: Gaussian setting time for solute transport in fluvial systems. Water Resources Research 47, W08601 (2011). [71] Zhang, Y, Meerschaert, M.M., Packman, A.I.: Linking fluvial bed sediment transport across scales. Geophysical Research Letters 39, L20404 (2012). 93