randomness and analysis a tutorial
play

Randomness and analysis: a tutorial Part I: Randomness notions and - PowerPoint PPT Presentation

Randomness and analysis: a tutorial Part I: Randomness notions and almost everywhere theorems Andr e Nies CCC 2015, Kochel am See 1/1 Di ff erentiability Di ff erentiability of a function f at a real z means that the rate of change


  1. The implications are proper (1) ) Martin-L¨ of random 6( computably random A set C is called high if ; 00  T C 0 . Equivalently, C computes a function that dominates each computable function (Martin, 1966). Theorem (N., Stephan, Terwijn, 2005) Every high set C Turing computes a set Z that is computably random. I Let h L e i e 2 N be a list of all partial computable martingales, I Define Z so that the martingale L = P e 2 � e L e is bounded along Z . I Use highness of C to deal with partiality. 19/1

  2. The implications are proper (2) On the other hand, if a computably enumerable set C is Turing above a random, then C is Turing equivalent to the halting problem ; 0 by the “Arslanov Completeness Critierion”. There is a high computably enumerable set C < T ; 0 . Therefore Martin-L¨ of random 6( computably random Another way to separate the ML and computable randomness: use the (prefix-free) Kolmogorov complexity of the initial segments. For ML-random Z we have K ( Z � n ) + O (1) � n . There is a computably random Y such that K ( Y � n ) = O (log n ). 20/1

  3. The implications are proper (3) ) computably random 6( Schnorr random. I First proved by Yongge Wang. I It is shown by a direct construction (see e.g. N’s book “Computability and Randomness”, Ch. 7). I N., Stephan, Terwijn, 2005 separate the two notions in each high degree. Note that any separation has to occur within the high degrees: Theorem (N., Stephan, Terwijn, 2005) If Z is not high and Schnorr random, then Z is ML-random. 21/1

  4. 3. E ff ective versions of almost everywhere theorems 22/1

  5. E ff ective almost everywhere theorems and randomness The “almost everywhere” theorems didn’t tell us whether the given object is well-behaved at a particular real. Now consider the case where the given object is algorithmic in some sense. I How strong an algorithmic randomness notion for a real z is needed to make the theorem hold at z ? I Will the theorem in fact characterize the randomness notion? Once this is settled, we can provide “concrete” examples of reals at which the nice behaviour occurs. For instance, Chaitin’s Ω is ML-random. 23/1

  6. Continuing the story of e ff ective a.e. theorems, after Bishop (1967) and Demuth (1975) Recall Birkho ff ’s 1939 theorem: Let ( X, µ, T ) be a measure preserving system, and let f : X ! R is measurable. For µ -almost every x , the limit as N ! 1 of the averages of f � T i ( x ) over 0  i < N , exists. I V’yugin, 1999 (TCS) shows that ML-randomness su ffi ces for the e ff ective Birkho ff theorem. (Note that T : ✓ X ! X only needs to be defined µ -a.e.) I He uses Bishop, Thm. 6 on page 236, which is closely related to his result on BV (Thm. 7). I Hoyrup, Rojas, Galatolo 2010-13 develop e ff ective ergodic theory. 24/1

  7. Schnorr randomness and L 1 -computability Pathak (2009), Pathak, Rojas, and Simpson (2012) proved an e ff ective version of another of Lebesgue’s theorem (but taking into account only the existence of limits, not the value). z 2 [0 , 1] d is Schnorr random , for every L 1 -computable function g : [0 , 1] d ! R , Z 1 lim g exists . λ ( B r ( z )) r ! 0 + B r ( z ) Implication ( also due to Freer, Kjos-Hanssen, N., Stephan. 25/1

  8. E ff ective form of the first Lebesgue theorem A function f : [ a, b ] ! R is of bounded variation if n � 1 X V ( f ) = sup | f ( t i +1 ) � f ( t i ) | < 1 , i =1 the sup taken over all collections t 1  t 2  . . .  t n in [ a, b ]. Theorem (Brattka, Miller, N; to appear in TAMS) Let f : [0 , 1] ! R be non-decreasing and computable. Then z is computably random ) f 0 ( z ) exists. I Under the weaker hypothesis that f has bounded variation, f 0 ( z ) exists for each Martin-L¨ of random real z , but not necessarily for each computably random. (Demuth, 1975; Brattka, Miller,N. ta). I Some depth here that doesn’t show in classical analysis. Jordan decomposition f = g 0 � g 1 for nondecreasing g i is not e ff ective! 26/1

  9. Functions-to-tests To prove the first result: If f is computable nondecreasing, we (uniformly in f ) build a computable martingale M such that f 0 ( z ) fails to exist ) M succeeds on z . (I will give detail later when I do the polynomial time computable case.) Corollary Each computable nondecreasing function f is di ff erentiable at a (uniformly obtained) computable real. PROOF: Each computable martingale fails on some computable real, which can be obtained uniformly. 27/1

  10. Converses (tests-to-functions) I Both the nondecreasing and the bounded variation cases also have converses: if z is not random in the appropriate sense, then some computable function of the respective type fails to be di ff erentiable at z (BMN, to appear). I So one could take the di ff erentiability properties for classes of e ff ective functions as definitions of randomness notions! z is computably random , each computable nondecreasing function is di ff erentiable at z z is Martin-L¨ of random , each computable function of bd. variation di ff erentiable at z . 28/1

  11. A new proof of Demuth’s result Here is the proof of Brattka/Miller/N. (TAMS, ta) of the result of Demuth on BV function. We get a stronger form: Let r be a Martin-L¨ of random real. Suppose f is uniformly computable on the rationals, and f is of bounded variation. Then f 0 ( r ) exists. I By Jordan’s result, f = h 0 � h 1 for some nondecreasing functions h 0 , h 1 . I One can show that r is Martin-L¨ of random (hence computably random) relative to some oracle set X encoding such a pair h 0 , h 1 . I By the previous theorem, relativized to X , the h i are both di ff erentiable at r . Thus f 0 ( r ) = h 0 0 ( r ) � h 0 1 ( r ) exists. ⇤ 29/1

  12. The strength of the Jordan decomposition theorem I Note that the pairs h 0 , h 1 with f = h 0 � h 1 (not necessarily continuous) can be seen as a Π 0 1 class P . I We obtain the decomposition because r is random in some member of P (the “low for r ” basis theorem). Results by Greenberg, N., Yokoyama, Slaman (see 2013 Logic Blog and upcoming project report by Marcus Triplett) show: I Jordan decomposition of any continuous BV f into continuous functions h 0 , h 1 is equivalent to ACA over RCA . I Jordan decomposition of any continuous BV f into nondecreasing functions h 0 , h 1 is equivalent to WKL over RCA . 30/1

  13. Randomness notions given by function classes (BMN ta) 31/1

  14. 4. Polynomial time randomness and di ff erentiability 32/1

  15. Special Cauchy names I A Cauchy name is a sequence of rationals ( p i ) i 2 N such that 8 k > i | p i � p k |  2 � i . I We represent a real x by a Cauchy name converging to x . For feasible analysis, we use a compact set of Cauchy names: the signed digit representation of a real. Such Cauchy names, called special, have the form i X b k 2 � k , p i = k =0 where b k 2 { � 1 , 0 , 1 } . (Also, b 0 = 0 , b 1 = 1.) So they are given by paths through { � 1 , 0 , 1 } ω , something a resource-bounded Turing machine can process. 33/1

  16. Polynomial time computable functions The following has been formulated in equivalent ways by Ker-i-Ko (1989), Weihrauch (2000), Braverman (2008). Definition A function g : [0 , 1] ! R is polynomial time computable if there is a polynomial time TM turning every special Cauchy name for x 2 [0 , 1] into a special Cauchy name for g ( x ). This means that the first n symbols of a special Cauchy name for g ( x ) can be computed in time polynomially in n , thereby using polynomially many symbols of the oracle tape that holds a special Cauchy name for x . 34/1

  17. Examples of polynomial time computable functions I Functions such as e x , sin x are polynomial time computable. I To see this one uses rapidly a converging approximation sequence, such as e x = P n x n /n !. I As Braverman (2008) points out, e x is computable in time O ( n 3 ). I Namely, from O ( n 3 ) symbols of x we can in time O ( n 3 ) compute an approximation of e x with error  2 � n . I Better algorithms may exist (e.g. search the 1987 book J. Borwein and P. Borwein, Pi and the AGM). I Breutzman, Juedes and Lutz (MLQ, 2004) have given an example of a polynomial time computable function that is no-where di ff erentiable. It is a variant of the Weierstrass function P n 2 � n cos(5 n π x ). 35/1

  18. Polynomial time randomness Recall that a betting strategy, or martingale, is a function M : 2 < ω ! R + 0 such that M ( σ 0) + M ( σ 1) = 2 M ( σ ) for each string σ . Definition A betting strategy M : 2 < ω ! R is called polynomial time computable if from a string σ and an i 2 N we can, in time polynomial in | σ | + i , compute the i -th component of a special Cauchy name for M ( σ ). In this case we can compute a polynomial time martingale in base 2 dominating M (Schnorr / Figueira-N). Definition We say Z 2 2 N is polynomial time random if no polynomial time betting strategy succeeds on Z . 36/1

  19. Polynomial time randomness Definition We say Z 2 2 N is polynomial time random if no polynomial time betting strategy succeeds on Z . I This was first studied in Yongge Wang’s 1992 thesis (Uni Heidelberg). I Figueira, N 2013 showed that the notion is base invariant: it is about reals rather than sequences of digits for a fixed base (such as 2). Proposition (Existence in super polynomial time classes) Suppose the function t ( n ) is time constructible and dominates all polynomials. There is polynomial random Z computable in time O ( t ( n )) (i.e. the language consisting of the initial segments of Z is O ( t )-computable). 37/1

  20. Lebesgue’s Thm (A) and its converse in the polytime setting Theorem (N., STACS 2014) The following are equivalent for a real z 2 [0 , 1]. (I) z (written in binary expansion) is polynomial time random (II) f 0 ( z ) exists, for each non-decreasing function f that is polynomial time computable. I The same method works for primitive recursive randomness/functions, and even computable randomness/computable functions. I So this also yields new proof of Brattka/Miller/Nies. 38/1

  21. Proof of the easy direction (II) ! (I) Suppose that f 0 ( z ) exists, for each non-decreasing function f that is polynomial time computable. We want to show that z is polynomial time random. Let S g ( σ ) denote the slope of a non-decreasing function g at the basic dyadic interval given by 1 string σ . This is a betting strategy. ) 1 1 Slope = M(1) ( M = e Essentially, each betting p o l S strategy M is of the form S g for nondecreasing g . If M is Slope = M(10) g polynomial time then so is g . 0.10 0.11 1.0 Since g 0 ( z ) exists, M is bounded along z . 39/1

  22. Slopes and their limits For a function f : ✓ R ! R , for a pair a, b of distinct reals let S f ( a, b ) = f ( a ) � f ( b ) . a � b For f defined on the rationals, the lower and upper (pseudo-)derivatives are D f ( x ) = lim inf h ! 0 + { S f ( a, b ) | a  x  b ^ 0 < b � a  h } , e e h ! 0 + { S f ( a, b ) | a  x  b ^ 0 < b � a  h } . Df ( x ) = lim sup where a, b range over rationals in [0 , 1]. Example: f ( x ) = x sin(1 /x ). f (0) = � 1 , e D Df (0) = 1. e For f defined in a nbhd of x and continuous at x , f ( x ) = e f 0 ( x ) exists i ff D Df ( x ) < 1 . e 40/1

  23. Slopes at basic dyadic intervals The subscript 2 indicates restriction to basic dyadic intervals [ σ ], where σ is a string, containing z : e { S f ( σ ) | x 2 [ σ ] } . D 2 f ( x ) = lim sup | σ | !1 Recall: if f is non-decreasing then M ( σ ) = S f ( σ ) is a betting strategy. We say that M converges on z if lim n M ( Z � n ) exists. We have the following basic connections: I M succeeds on z , e D 2 f ( z ) = 1 . 2 f ( z ) = e I M converges on z , D D 2 f ( z ) < 1 e 41/1

  24. Proof of the harder direction (I) ! (II) Now suppose that z = 0 .Z 2 [0 , 1] is polynomial time random. We want to show that f 0 ( z ) exists, for each non-decreasing function f that is polynomial time computable. I Consider the polynomial time computable betting strategy M ( σ ) = S f ( σ ) . I lim n M ( Z � n ) exists and is finite for each polynomially random Z . This is an e ffi cient version of Doob’s martingale convergence theorem. 2 f ( z ) = e I Therefore D D 2 f ( z ) < 1 . e 42/1

  25. Porosity Assume for a contradiction that f 0 ( z ) fails to exist. We have oscillation of slopes of f at arbitrarily small intervals around z . We want success of a betting strategy at basic dyadic intervals corresponding to prefixes of Z . I First suppose that e D 2 f ( z ) < p < e Df ( z ). D 2 f ( z ) < p there is a string σ ⇤ � Z such that I Since e 8 σ [ Z � σ ⌫ σ ⇤ ) S f ( σ )  p ]. I Choose k with p (1 + 2 � k +1 ) < e Df ( z ). Let � denote the prefix relation of strings. The next lemma says that [ σ ⇤ ] � S { ( σ ): σ ⌫ σ ⇤ ^ S f ( σ ) > p } is porous at z . Lemma (High slopes at dyadic intervals) There are arbitrarily large n such that S f ( τ n ) > p for some basic dyadic interval [ τ n ] of length 2 � n � k which is contained in [ z � 2 � n +2 , z + 2 � n +2 ]. 43/1

  26. We may suppose σ ⇤ is the empty string, i.e., S f ( σ )  p for all dyadic intervals [ σ ] containing z . By the lemma, there are arbitrarily large n such that S f ( τ n ) > p for some basic dyadic interval [ τ n ] of length 2 � n � k which is contained in [ z � 2 � n +2 , z + 2 � n +2 ]. Good case: there are infinitely many n with η = Z � n � 4 � τ n . Then the strategy that from such η on bets everything on the strings of length n + k other than τ n gains a fixed factor 2 k +4 / (2 k +4 � 1) on Z each time. Also, it never goes down on Z , so it succeeds. Bad case: for almost all n we have Z � n � 4 6� τ n . This means 0 . τ n is on the left side of z . So the strategy can’t use τ n as it splits o ff from Z before η is read. 44/1

  27. The shifting–by–1/3 trick Fix m 2 N . For k 2 Z consider an interval I = [ k 2 � m , ( k + 1)2 � m ] . For r 2 Z consider an interval J = 1 / 3 + [ r 2 � m , ( r + 1)2 � m ] . The distance between an endpoint of I and an endpoint of J is at least 1 / (3 · 2 m ). To see this: assume that k 2 � m � ( p 2 � m + 1 / 3) < 1 / (3 · 2 m ). This yields (3 k � 3 p � 2 m ) / (3 · 2 m ) < 1 / (3 · 2 m ), and hence 3 | 2 m , a contradiction. 45/1

  28. Using this trick to finish the proof of (I) ! (II) We may assume that z > 1 / 2. In the “bad” case that Z � n � 4 6� τ n for almost all τ n , we instead bet on the dyadic expansion Y of z � 1 / 3. I Given η 0 = Y � n � 4 , look for an extension τ 0 � η 0 of length n + k + 1, such that 1 / 3 + [ τ 0 ] ✓ [ τ ] for a string [ τ ] with S f ( τ ) > p . (Then Y 62 [ τ 0 ].) I If it is found, bet everything on the other extensions of η 0 of that length n + k + 1. This strategy gains a fixed factor 2 k +5 / (2 k +5 � 1) on Y each time n is as above. It never goes down on Y , so it succeeds. So we get a polytime martingale that wins on z � 1 / 3. By Figueira and N (2013), polytime randomness is base invariant, so z � 1 / 3 is polynomially random. This yields a contradiction. The case D f ( z ) < D 2 f ( z ) is analogous, using a “low dyadic e e slopes” lemma instead. 46/1

  29. Shifted dyadic versus full di ff erentiability For a rational q let D q be the collection of intervals of the form q + [ k 2 � m , ( k + 1)2 � m ] where k 2 Z , m 2 N . Question Let f : [0 , 1] ! R be continuous nondecreasing, and let z 2 (0 , 1). Suppose that for each rational q , [ a,b ] 2 D q , z 2 [ a,b ] , b � a ! 0 S f ( a, b ) lim exists. Is f already di ff erentiable at z ? 47/1

  30. 5. Di ff erentiability of Lipschitz functions 48/1

  31. Computable randomness and Lipschitz functions Recall that f is Lipschitz if | f ( x ) � f ( y |  C ( | x � y | ) for some C 2 N . Theorem (Freer, Kjos, N, Stephan, Computability, 2014) A real z is computably random ( ) each computable Lipschitz function f : [0 , 1] ! R is di ff erentiable at z . ) : Write f ( x ) = ( f ( x ) + Cx ) � Cx . Then f ( x ) + Cx is = computable and non-decreasing. From the monotone case (BMN), we obtain a test (martingale) for this function. If f 0 ( z ) does not exists, then z fails this test. ( = : Turn success of a martingale on a real into oscillation of the slopes, around the real, of a Lipschitz function. 49/1

  32. Rademacher’s theorem Theorem (Rademacher, 1920) Let f : [0 , 1] n ! R be Lipschitz. Then the derivative Df ( z ) (an element of R n ) exists for almost every vector z 2 [0 , 1] n . To define computable randomness of a vector z 2 [0 , 1] n : I Take the binary expansion of the n components of z . I We can bet on the corresponding sequence of blocks of n bits. Rute (2012) studies this notion, for instance invariance under computable measure preserving operators. 50/1

  33. E ff ective form of Rademacher Theorem (Galicki and Turetsky, arxiv.org/abs/1410.8578 ) z 2 [0 , 1] n is computably random ) every computable Lipschitz function f : [0 , 1] n ! R is di ff erentiable at z . For a vector v , the directional derivative Df ( z ; v ) is the derivative of the function t ! f ( z + tv ) at 0. The proof has three steps: I all partial derivatives exist at z I all directional derivatives for computable directions exist I Df ( z ; v ) is linear on computable directions v Since f is Lipschitz this show that f Gˆ ateaux-di ff erentiable at z : all directional derivatives exist, and the value is linear in the direction. Again since f is Lipschitz, this yields the full di ff erentiability of f at z . ⇤ 51/1

  34. Other approaches to e ff ective Lipschitz functions The polynomial time case is open. Question Suppose z 2 [0 , 1] n is polynomially random. Is every polynomial time Lipschitz function f : [0 , 1] n ! R is di ff erentiable at z ? I Abbas Edalat has developed an approach to di ff erentiability of e ff ective Lipschitz functions using domain theory. See his recent paper in TCS. I It involves the Clarke gradient (a set-valued derivative) to get around the measure 0 set where the function is not classically di ff erentiable. 52/1

  35. 6. Two more almost everywhere results: Carleson-Hunt (1966/68) and Weyl (1916) 53/1

  36. Carleson-Hunt Thm (suggested by Manfred Sauter) Theorem (Carleson, 1966 for p = 2; improved by Hunt 1968) Let f 2 L p [ � π , π ] be a periodic function. Then the Fourier series c N ( z ) = P f ( n ) e inz converges for almost every z . | n |  N ˆ Question Suppose f is L p -computable for computable p > 1. Which randomness property of z su ffi ces to make the sequence c N ( z ) converge? I We say z is weakly 2-random if z is in no null Π 0 2 set. This properly implies Martin-L¨ of randomness. I As an easy consequence of Carleson-Hunt theorem, weak 2-randomness of z su ffi ces. For fixed rationals α < β , the statement that, say, Re c N ( z ) oscillates between values < α and > β is Π 0 2 . 54/1

  37. E ff ective Weyl Theorem Theorem (Weyl, 1916) Let ( a i ) i 2 N be a sequence of distinct integers. Then for almost every real z , the sequence a i z mod 1 is uniformly distributed in [0 , 1]. Suppose now ( a i ) i 2 N is computable. Avigad (2012) shows that I Schnorr randomness of z su ffi ces to make the conclusion of Weyl’s theorem hold. I There is a z satisfying the conclusion of the theorem which is in some null e ff ectively closed set (hence not even “Kurtz random”). 55/1

  38. 7. E ff ective ergodic theory: multiple recurrence 56/1

  39. Classical theory A measurable operator T on a probability space ( X, B , µ ) is measure preserving if µT � 1 ( A ) = µA for each A 2 B . The following is Furstenberg’s multiple recurrence theorem (1977); see Furstenberg’s book on recurrence, 2014 edition, Thm. 7.15. Theorem Let ( X, B , µ ) be a probability space. Let T 1 , . . . , T k be commuting measure preserving operators on X . For each P 2 B with µP > 0, there is n > 0 such that µ ( T i T � n ( P )) > 0. i With a little measure theory one can easily strengthen this to an “almost-everywhere” type result: a.e. z 2 P 9 n [ z 2 T i T � n ( P )]. i 57/1

  40. k -recurrence in Cantor space Let X = 2 N with the shift operator S : X ! X that takes the first bit o ff a sequence. Definition Let P ✓ 2 N be measurable, and Z 2 2 N . We say that Z is k -recurrent in P if S n ( Z ) , S 2 n ( Z ) , . . . , S kn ( Z ) 2 P for some n � 1, i.e. Z 2 T 1  i  k S � ni ( P ). Theorem (Downey, Nandakumar, N., in preparation) Let P ✓ 2 N be a Π 0 1 class of positive measure. Each Martin-L¨ of random Z is k -recurrent in P , for each k � 1. Martin-L¨ of-randomness is necessary even for k = 1. If Z is not ML-random, no “tail” S n ( Z ) is in the Π 0 1 class of positive measure P = { Y : 8 r K ( Y � r ) � r � 1 } by the Levin-Schnorr Theorem. 58/1

  41. General Conjecture It is likely that an e ff ective multiple recurrence theorem holds in full generality for ML-randomness and Π 0 1 sets. Conjecture Let ( X, µ ) be a computable probability space. Let T 1 , . . . , T n be computable measure preserving transformation that commute pairwise. Let P be a Π 0 1 class with µP > 0. If Z 2 P is ML-random then 9 n V i T n i z 2 P . I By the classical of Furstenberg, this holds for weakly 2-random Z (i.e., in no null Π 0 2 class). I Jason Rute has pointed out that if µ P is computable, then Schnorr randomness of Z is su ffi cient, also by the classical result. A draft of this work is available on the 2015 Logic Blog. 59/1

  42. Randomness and analysis: a tutorial Part II: Lebesgue density and its applications to randomness Andr´ e Nies CCC 2015, Kochel am See 1/24

  43. Density Let � denote uniform (Lebesgue) measure. Definition Let E be a subset of [0 , 1]. The (lower) density of E at a real z is � ( J \ E ) ⇢ ( E | z ) = lim inf , | J | z 2 J , | J | ! 0 where J ranges over intervals. This gauges how much, at least, of E is in intervals that zoom in on z . ⇢ ( E | z ) is the limit over intervals containing z . Clearly ⇢ ( E | z ) = 1 $ ⇢ ( E | z ) = 1. 2/24

  44. Lebesgue’s Theorem: towards an e ff ective version Recall: ⇢ ( E | z ) = lim inf J interval ,z 2 J, | J | ! 0 � ( J \ E ) / | J | . Theorem (Lebesgue Density Theorem, 1910) Let E ✓ [0 , 1] be measurable. Then for almost every z 2 [0 , 1] : if z 2 E , then ⇢ ( E | z ) = 1 . I For open E , this is immediate, and actually holds for all z 2 [0 , 1]. I For closed E , this is the simplest case where there is something to prove. E ✓ [0 , 1] is e ff ectively closed (or Π 0 1 ) if there is an e ff ective list of open intervals with rational endpoints that has union [0 , 1] \ E . Definition (Main) We say that a real z is a density-one point if ⇢ ( E | z ) = 1 for every e ff ectively closed E 3 z . 3/24

  45. Martin-L¨ of randomness and density Does Martin-L¨ of randomness ensure that an e ff ectively closed E ✓ [0 , 1] with z 2 E has density one at z ? Answer: NO! Example I Let E 6 = ; , E ✓ [0 , 1] be an e ff ectively closed set containing only Martin-L¨ of randoms. I E.g., E = [0 , 1] \ S 1 where h S r i r 2 N is a universal ML-test. I Let z = min( E ). I Then ⇢ ( E | z ) = 0 even though z is ML-random. 4/24

  46. Density randomness Definition Let E be a measurable subset of 2 N . The lower dyadic density of E at Z 2 2 N is n !1 2 n � ([ Z � n ] \ E ) . ⇢ 2 ( E | Z ) = lim inf Definition We say that Z ✓ N is density random if Z is ML-random and ⇢ 2 ( P | Z ) = 1 for each Π 0 1 class P 2 Z . For ML-random Z , one can equivalently require the full density equals 1 in the setting of reals, by a result of Khan and Miller (2013). 5/24

  47. Three characterisations of density randomness Theorem The following are equivalent for Z 2 2 N , z = 0 .Z . I Z is density random I [Madison group, 2012] Each left-c.e. martingale M converges: lim n M ( Z � n ) exists ( M is left-c.e. if M ( � ) is a left-c.e. real uniformly in string � ) I [N., 2014] g 0 ( z ) exists for each interval-c.e. function g I [Miyabe, N., Zhang 2013] For each integrable lower + , the “averaging” semicomputable function f : [0 , 1] ! R statement of the Lebesgue di ff erentiation theorem holds at z . For background and complete proofs see Miyabe, N., Zhang 2013. The continuous interval-c.e. functions with g (0) = 0 are precisely the variation functions of computable functions by Freer et al. 2014. 6/24

  48. 2. Anti-random sequences 000000000000000000000000000000000000000000000000000000 000000000000000000000000000000000000000000000000000000 000000000000000000000001000000000000000001111111111111 111111111111111111111111111111111111111111111100000000 000000000000000000111000000000000000000000000000000000 . . . 7/24

  49. Basic objects of computability theory 8/24

  50. Basic objects of computability theory I The computable subsets of N Computable sets 8/24

  51. Basic objects of computability theory ∅ ' I The computable subsets of N I the halting problem ; 0 Computable sets 8/24

  52. Basic objects of computability theory ∅ ' I The computable subsets of N I the halting problem ; 0 ∆ 0 !!! sets 2 I Turing reducibility  T I the ∆ 0 2 sets ( A  T ; 0 ) Computable sets 8/24

  53. Basic objects of computability theory ∅ ' I The computable subsets of N I the halting problem ; 0 C.e. sets I Turing reducibility  T I the computably enumerable sets Computable sets 8/24

  54. Adding the world of (anti-)randomness Computable sets 9/24

  55. Adding the world of (anti-)randomness Ω Z I The Martin-L¨ of random sets Z , such as Chaitin’s halting probability Ω . We have Ω ⌘ T ; 0 . Computable sets 9/24

  56. Adding the world of (anti-)randomness Ω I The antirandom ( K -trivial) sets. K-trivial sets 9/24

  57. Adding the world of (anti-)randomness Ω I The antirandom ( K -trivial) sets. D (c.e.) If A is K -trivial, then there is c.e. A K -trivial set D � tt A . (N. 2005) K-trivial sets 9/24

  58. Kuˇ cera’s theorem and the covering problem Computable sets 10/24

  59. Kuˇ cera’s theorem and the covering problem Let Z be a random ∆ 0 2 set. Then there is a c.e., incomputable set A  T Z . (Kuˇ cera, 1986) Z C.e. sets A Computable sets 10/24

  60. Kuˇ cera’s theorem and the covering problem Let Z be random with Z 6� T ; 0 . ∅ � Let A  T Z be c.e. Z Then A is K -trivial. (Hirschfeldt, N., Stephan, 2007) C.e. sets A K-trivial sets 10/24

  61. Kuˇ cera’s theorem and the covering problem Let Z be random with Z 6� T ; 0 . ∅ � Let A  T Z be c.e. Then A is K -trivial. (Hirschfeldt, N., Stephan, 2007) Covering problem (Stephan, 2004) C.e. sets Let A be a c.e. K -trivial set. A K-trivial sets 10/24

  62. Kuˇ cera’s theorem and the covering problem Let Z be random with Z 6� T ; 0 . ∅ � Let A  T Z be c.e. Z Then A is K -trivial. (Hirschfeldt, N., Stephan, 2007) Covering problem (Stephan, 2004) C.e. sets Let A be a c.e. K -trivial set. Is there a ML-random Z � T A A with Z 6� T ; 0 ? K-trivial sets 10/24

  63. Kuˇ cera’s theorem and the covering problem Let Z be random with Z 6� T ; 0 . ∅ � Let A  T Z be c.e. Z Then A is K -trivial. (Hirschfeldt, N., Stephan, 2007) Covering problem (Stephan, 2004) C.e. sets Let A be a c.e. K -trivial set. Is there a ML-random Z � T A A with Z 6� T ; 0 ? We may omit the assumption that A is c.e.: if not, replace A by a c.e. K -trivial set D above A . 10/24

  64. A strong solution to the covering problem Theorem 5 + 2 authors ∅ � There is a ML-random set Z < T ; 0 above all the K -trivials. Z K-trivial sets 11/24

  65. A strong solution to the covering problem Theorem 5 + 2 authors ∅ � There is a ML-random set Z < T ; 0 above all the K -trivials. Z How random can Z be? K-trivial sets 11/24

  66. A strong solution to the covering problem Theorem 5 + 2 authors ∅ � There is a ML-random set Z < T ; 0 above all the K -trivials. Z How random can Z be? Answer: not much more than Martin-L¨ of random. K-trivial sets 11/24

  67. A strong solution to the covering problem Theorem 5 + 2 authors ∅ � There is a ML-random set Z < T ; 0 above all the K -trivials. Z How random can Z be? Answer: not much more than Martin-L¨ of random. How close to ; 0 must Z lie? K-trivial sets 11/24

  68. A strong solution to the covering problem Theorem 5 + 2 authors ∅ � There is a ML-random set Z < T ; 0 Z above all the K -trivials. How random can Z be? Answer: not much more than Martin-L¨ of random. How close to ; 0 must Z lie? Answer: Z is very close to ; 0 . K-trivial sets 11/24

  69. Background on antirandom sets 12/24

  70. Descriptive string complexity K Consider a partial computable function from binary strings to binary strings (called machine). It is called prefix-free if its domain is an antichain under the prefix relation of strings. There is a universal prefix-free machine U : for every prefix-free machine M , M ( � ) = y implies U ( ⌧ ) = y for some ⌧ with | ⌧ |  | � | + d M , and the constant d M only depends on M . The prefix-free Kolmogorov complexity of string y is the length of a shortest U -description of y : K ( y ) = min {| � | : U ( � ) = y } . 13/24

  71. Definition of K -triviality In the following, we identify a natural number n with its binary representation (as a string). For a string ⌧ , up to additive const we have K ( | ⌧ | )  K ( ⌧ ), since we can compute | ⌧ | from ⌧ . 14/24

  72. Definition of K -triviality In the following, we identify a natural number n with its binary representation (as a string). For a string ⌧ , up to additive const we have K ( | ⌧ | )  K ( ⌧ ), since we can compute | ⌧ | from ⌧ . Definition (going back to Chaitin, 1975) An infinite sequence of bits A is K -trivial if, for some b 2 N , 8 n [ K ( A � n )  K ( n ) + b ], namely, all its initial segments have minimal K -complexity. 14/24

  73. Definition of K -triviality In the following, we identify a natural number n with its binary representation (as a string). For a string ⌧ , up to additive const we have K ( | ⌧ | )  K ( ⌧ ), since we can compute | ⌧ | from ⌧ . Definition (going back to Chaitin, 1975) An infinite sequence of bits A is K -trivial if, for some b 2 N , 8 n [ K ( A � n )  K ( n ) + b ], namely, all its initial segments have minimal K -complexity. It is not hard to see that K ( n )  2 log 2 n + O (1). 14/24

  74. Definition of K -triviality In the following, we identify a natural number n with its binary representation (as a string). For a string ⌧ , up to additive const we have K ( | ⌧ | )  K ( ⌧ ), since we can compute | ⌧ | from ⌧ . Definition (going back to Chaitin, 1975) An infinite sequence of bits A is K -trivial if, for some b 2 N , 8 n [ K ( A � n )  K ( n ) + b ], namely, all its initial segments have minimal K -complexity. It is not hard to see that K ( n )  2 log 2 n + O (1). Z is random , 8 n [ K ( Z � n ) > n � O (1)] 14/24

  75. Definition of K -triviality In the following, we identify a natural number n with its binary representation (as a string). For a string ⌧ , up to additive const we have K ( | ⌧ | )  K ( ⌧ ), since we can compute | ⌧ | from ⌧ . Definition (going back to Chaitin, 1975) An infinite sequence of bits A is K -trivial if, for some b 2 N , 8 n [ K ( A � n )  K ( n ) + b ], namely, all its initial segments have minimal K -complexity. It is not hard to see that K ( n )  2 log 2 n + O (1). A is K -trivial , 8 n [ K ( A � n )  K ( n ) + O (1)] 14/24

  76. Definition of K -triviality In the following, we identify a natural number n with its binary representation (as a string). For a string ⌧ , up to additive const we have K ( | ⌧ | )  K ( ⌧ ), since we can compute | ⌧ | from ⌧ . Definition (going back to Chaitin, 1975) An infinite sequence of bits A is K -trivial if, for some b 2 N , 8 n [ K ( A � n )  K ( n ) + b ], namely, all its initial segments have minimal K -complexity. It is not hard to see that K ( n )  2 log 2 n + O (1). Z is random , 8 n [ K ( Z � n ) > n � O (1)] A is K -trivial , 8 n [ K ( A � n )  K ( n ) + O (1)] Thus, being K -trivial means being far from random . 14/24

  77. Connecting density and K -triviality This is based on the following work: [Oberwolfach] Bienvenu, Greenberg, Kuˇ cera, N. Turetsky 2012 JEMS, to appear [Berkeley] Day and Miller 2012 Math. Research Letters, to appear [Paris] Bienvenu, Miller, H¨ olzl and N. 2011 STACS 2012, JML 2014 15/24

  78. Turing incompleteness and positive density Definition We say that a real z is a positive density point if ⇢ ( E | z ) > 0 for every e ff ectively closed E 3 z . For a real z 62 Q , let Z 2 2 N denote its binary expansion: z = 0 .Z . Theorem (Paris) Let z be a Martin-L¨ of random real. Then Z is NOT above the halting problem ; 0 , z is a positive density point. 16/24

  79. The main connection of density and K -trivials Recall: ⇢ ( E | z ) = lim inf | J | → 0 ,z ∈ J � ( J \ E ) / | J | . Definition (Recall) We say that a real z is a density-one point if ⇢ ( E | z ) = 1 for every e ff ectively closed E 3 z . In other words, z satisfies the Lebesgue Theorem for e ff ectively closed sets. Z ∅ � Theorem (Oberwolfach) Let z be a Martin-L¨ of random real. Suppose z is NOT a density-one point. Then Z is above all the K -trivials. K-trivial sets 17/24

  80. The main connection of density and K -trivials Z ∅ � Theorem Let z be a Martin-L¨ of random real. Suppose z is not a density-one point. Then Z is above all the K -trivials. K-trivial sets 18/24

  81. The main connection of density and K -trivials Z ∅ � Theorem Let z be a Martin-L¨ of random real. Suppose z is not a density-one point. Then Z is above all the K -trivials. K-trivial sets To solve the covering problem, we need to know: Does Z as in the picture exist? 18/24

  82. The main connection of density and K -trivials Z ∅ � Theorem Let z be a Martin-L¨ of random real. Suppose z is not a density-one point. Then Z is above all the K -trivials. K-trivial sets To solve the covering problem, we need to know: Does Z as in the picture exist? Where do we get a ML-random set Z 6� T ; 0 that is not a density-one point? 18/24

Download Presentation
Download Policy: The content available on the website is offered to you 'AS IS' for your personal information and use only. It cannot be commercialized, licensed, or distributed on other websites without prior consent from the author. To download a presentation, simply click this link. If you encounter any difficulties during the download process, it's possible that the publisher has removed the file from their server.

Recommend


More recommend