REFINEMENTS OF GAL’S THEOREM AND APPLICATIONS

3y ago
10 Views
2 Downloads
350.46 KB
16 Pages
Last View : 1m ago
Last Download : 3m ago
Upload by : Wren Viola
Transcription

REFINEMENTS OF GÁL’S THEOREM AND APPLICATIONSMARK LEWKO AND MAKSYM RADZIWILLAbstract. We give a simple proof of a well-known theorem of Gál and of therecent related results of Aistleitner, Berkes and Seip [1] regarding the size of GCDsums. In fact, our method obtains the asymptotically sharp constant in Gál’stheorem, which is new. Our approach also gives a transparent explanation of therelationship between the maximal size of the Riemann zeta function on vertical linesand bounds on GCD sums; a point which was previously unclear. Furthermorewe obtain sharp bounds on the spectral norm of GCD matrices which settles aquestion raised in [2]. We use bounds for the spectral norm to show that seriesformed out of dilates of periodic functions of bounded variation converge almosteverywhere if the coefficients of the series are in L2 (log log 1/L)γ , with γ 2. Thiswas previously known with γ 4, and is known to fail for γ 2. We also developa sharp Carleson-Hunt-type theorem for functions of bounded variations whichsettles another question raised in [1]. Finally we obtain almost sure bounds forpartial sums of dilates of periodic functions of bounded variations improving [1].This implies almost sure bounds for the discrepancy of {nk x} with nk an arbitrarygrowing sequences of integers.1. IntroductionLet 1 n1 n2 . . . nk be an arbitrary sequence of integers. The problem ofbounding GCD sums1 X (ni , nj )2α, 0 α 1(1)k i,j k (ni nj )αarises naturally in metric Diophantine approximation, following Koksma’s initialwork [19] (see also [10]).While estimating (1) for many specific sequences is straightforward, the problemof determining the maximal size of (1) among all sequences with k terms is muchmore subtle and the case α 1 was posed as a prize problem by the Scientific Societyat Amsterdam in 1947 on Erdös’s suggestion.Date: September 18, 2014.2010 Mathematics Subject Classification. 11C20, 42A20, 42A61, 42B05.The first author is supported by a NSF postdoctoral fellowship, DMS-12042 and the IAS Fundfor Mathematics, the second author was partially supported by NSF grant DMS-1001068.1

2MARK LEWKO AND MAKSYM RADZIWILLThe problem for α 1 was solved by I. S. Gál in 1949 [13]. Gál’s proof is a difficultcombinatorial argument spanning 20 pages. Gál showed that for α 1 the GCDsum (1) is bounded by C(log log k)2 with C 0 an absolute constant and moreoverthat this bound is optimal up to the value of the constant C 0.Our first contribution is a simple, two page proof, of Gál’s theorem. In additionour proof determines the optimal constant C as k , which is new.Theorem 1. As k ,sup1 n1 . nkX (ni , nj )26e2γ1 .k(log log k)2 i,j k ni njπ2The extremal sequence in Theorem 1 is supported on very smooth integers. Thegeneralization of (1) to 12 α 1 was studied by Dyer and Harman [10] whowere also interested in applications to metric diophantine approximations. Recently,Aistleitner, Berkes and Seip [1] showed that for 1/2 α 1, the GCD sum (1) isbounded by (log k)1 α .(2)exp C(α) ·(log log k)αThis bound is sharp up to the value of the constant C(α). Several authors haveremarked on the similarity between this estimate and the conjectured maximal sizeof the zeta function along a vertical line (see [21, 23]) which states, for 1/2 α 1,(3)sup log ζ(α it) t k(log k)1 α.(log log k)αOur method generalizes to the case 21 α 1 and allows us to rederive in a simpleway the results of Aistleitner, Berkes and Seip [1]. In addition, the proof we giveshows that the GCD sum (1) is essentially majorized by the square of the supremumof a certain random model of the zeta function, which is in fact used to conjecture(3) (see [21, 23] and Section 9 at the end of this paper). In particular the similaritybetween (2) and (3) is not a coincidence.Our method also generalizes to the spectral norm case, which is a key ingredientin the applications described later on. In particular, if we let c (c(1), . . . , c(n)),then one is interested in bounding the quantity1 X (ni , nj )2αsup· c(ni )c(nj ).αkck22 1 k i,j k (ni nj )Sharp bounds for the spectral norm have been established for 1/2 α 1 in [1,Theorem 5] but the case α 1 had remained open (see [1], [2]). Our main theoremsettles this problem.

GCD SUMS AND APPLICATIONS3Theorem 2. Let c (c(1), . . . , c(n)). Then, for α 1, 6e2γ 1 X (ni , nj )2α) (4)sup·c(n)c(n o(1)· (log log k)2 .jiα22k(nn)πi jkck2 1i,j kIn addition for 0 α 1 we re-derive the bounds obtained for the spectral norm byAistleitner, Berkes and Seip in [1, Theorem 5]. Precisely, for 21 α 1 we boundthe spectral norm by (log k)1 α exp 2C(α)(log log k)αand for 0 α 21 by p1 2αk· exp 2C(α) log k log log kwith C(α) an absolute constants depending only on α.The bound (4) answers a question raised by Aistleitner, Berkes, Seip and Weber[2] regarding the correct power of log log k for the spectral norm at α 1, where itis described as a “a profound problem” (see remarks after (33) in [2]). We note thatinequality (4) also has an asymptotically sharp constant. For example, Hilberdink[17] showed that when ni i the bound (4) is attained as k for a certain choiceof the coefficients c(k) (this however is not true for 1/2 α 1, see [2]). The boundis also attained when nk is choosen to be the extremal sequence in Theorem 1 andc(k) k 1/2 .The main idea in the proof of Theorem 2 is that one can write (1) as an integralinvolving the Riemann zeta-function (or, more precisely, a random model for theRiemann zeta-function) and then appeal to known distributional estimates for thisquantity.We note that while our method recovers completely Theorem 1 from [1] and goesbeyond when α 1, the bound for α 1/2 is nonetheless not optimal. It has beenconjectured in [2], correcting an older conjecture of Harman [16], that the optimalbound in the limiting case α 1/2 iss!log kexp C ·.log log kThe best results towards this conjecture are due to Bondarenko and Seip (see [5]),who come within a triple logarithm of the conjecture.We now focus on applications of Theorem 2, in the spirit of [1]. Let f be a functionsuch thatZ 1(5)f (x 1) f (x) ,f (x)dx 0.0

4MARK LEWKO AND MAKSYM RADZIWILLWe are interested in L2 conditions for the almost everywhere convergence of X(6)c f (n x) 1for 1 n1 n2 . . . an arbitrary sequence of integers, and in almost sure boundsforX(7)f (ni x).i kFrom the point of view of applications to metrical diophantine approximation anatural choice of f is f (x) {x} 21 , a function of bounded variation (which impliesthat the j-th Fourier coefficient of f is O(1/j)). Choosing f (x) χI ({x}) I in(7), with χI the indicator function of an interval I [0, 1] relates the problem ofbounding (7) to that of obtaining almost sure bounds for the discrepancy of {nk x}.For smooth functions such as f (x) sin 2πx, one can use a deep result of Carleson[6] to show that if ck is in 2 then (6) converges almost everywhere. However alreadyfor f (x) {x} 12 the condition ck 2 is insufficient. In [1, Section 6] it is shownthat for f (x) {x} 12 and any γ 2 there exists an increasing sequence of positiveintegers ni and a sequence of real numbers c , with X(8)c2 · (log log )γ 1for which (6) diverges almost everywhere.Aistleitner, Berkes and Seip [1, Theorem 3] complemented this negative result byshowing that if (8) holds for γ 4 then (6) converges almost everywhere for any fof bounded variation (and thus also f (x) {x} 21 ). In the theorem below we closethe remaining gap.Corollary 3. Let f be a function of bounded variation satisfying (5) . Let ck be asequence of real numbers such thatXc2 · (log log )γ 3for some γ 2. Then for every increasing sequence (n ) 1 the seriesXc f (n x) 1converges almost everywhere.This result is optimal in the sense that the exponent γ cannot be lowered anyfurther. We note that our coundition is roughly equivalent to c L2 · (log log 1/L)γ ,

GCD SUMS AND APPLICATIONS5γ 2, since the series composed of integers with c 1/ 2 converges absolutely.Our method of proof also allows us to recover the recent results in [2] obtained forfunctions f with Fourier coefficients decaying at a rate of j α with 1/2 α 1.Corollary 3 also improves a very recent result of Weber [25] where the same conclusionis obtained with a (log log k)4 /(log log log k)2 in place of (log log k)2 ε for the specialcase nk k.We also obtain the following improvement of a result in [1, Theorem 2].Corollary 4. Let f be a function of bounded variation satisfying (5). Then, foralmost every x,Xpf (n x) N log N (log log N )3/2 ε . NThis improves the exponent 52 obtained in [1, Theorem 2] to 23 . The optimalexponent is conjecturedto be 12 . The problem of obtaining almost everywhere boundsPon the quantity N f (n x) has a long history, partly motivated by the problemof obtaining almost sure bounds for the discrepancy of the sequence {nk x}. Indeedweaker estimates on this quantity (or special cases thereof) were obtained by Gál[13] (1949), Erdös and Koksma [11] (1949), Gál and Koksma [14] (1950), Cassels [7](1950), R. C. Baker [3] (1981), Aistleitner, Mayer and Ziegler (2010), and Aistleitner,Berkes and Seip [1] (2012). See also [2], [4], and [25].The key estimate in the proofs of Corollary 3 and Corollary 4 is an optimalCarleson-Hunt-type inequality for systems of dilated functions {f (n x)} with f ofbounded variation. The theorem below answers a question in [1] regarding the optimal version of the Carleson-Hunt theorem, in this setting (see remarks after Lemma4 in [1]). It would be fitting to call this a maximal analogue of Gál’s theorem.Theorem 5. Let f : T C be a complex-valued function on the circle with Fouriercoefficients, a(j), satisfying the decay condition a(j) O( j 1 ). Let n1 , n2 , . . . , nNbe a strictly increasing sequence of positive integers and c(k) a sequence of complexnumbers. Then,!2ZMNXX2maxc(k)f (nk x)(log log N ) c(k) 2 .T1 M Nk 1k 1This inequality with an exponent of 4 instead of 2 was obtained in [1, Lemma 4],it’s also shown there that the function (log log N )2 cannot be replaced by any slowergrowing function.There are several innovations compared to the proof in [1]. First, we perform asplitting according to the largest prime divisor, secondly, we use a majorant principleto handle the tails composed of large primes, and finally we use the ideas that enter

6MARK LEWKO AND MAKSYM RADZIWILLin our proof of Theorem 2 to handle the contribution of the large primes after theapplication of the majorant principle.Because of the splitting, which is done according to the largest prime factor, itis tempting to investigate if there are any links with the P-summation method ofFouvry and Tenenbaum, applied to trigonometric series as in de la Bretèche andTenenbaum’s paper [9].2. NotationWe use the usual asymptotic notation. For instance, we write X Y to indicatethat there exists a universal constant C such that X C Y . We let T denote theunit circle, and e(x) : e2πix for x T. For f L1 (T) we define the j-th Fouriercoefficient by the relationZc(j) : f (x)e( jx)dx.TWe let ζ(s) denote the Riemann zeta function.3. The random modelLet X(p) be a sequence of independent random variable, one for each prime p, andequidistributed on the torus T. For an integer n we letYX(n) : X(p)α .pα knThe random model of the zeta-function that we will be working with is the followingY X(p) 1ζ(σ, X) : 1 σ.ppNote that the product is convergent almost surely for σ 12 by Kolmogorov’s threeseries theorem. Note also that in an Lp sense (with p 0) for 21 σ 1,X X(n)ζ(σ, X) nσnand that(1E[X(n)X(m)] 0if n m.otherwiseInstead of working with the probabilistic model we could also work with the zetafunction itself, since for example,Z Thi1lim ζ(σ it) 2 · n it dt E ζ(σ, X) 2 · X(n) .T 2T T

GCD SUMS AND APPLICATIONS7To re-inforce this point, the distributional estimates that we use in Lemma 7 beloware known unconditionally for ζ(σ it) (see [21, 15, 24]). These results however areoften obtained by first passing to the random model, for this reason we did not seethe advantage of working with ζ(σ it) directly which relies on deeper machinery (forexample the zeros of ζ(s) enter the analysis). The random model described aboveis commonly used in the study of the Riemann zeta-function, we refer the reader to[21, 15, 24, 20, 8]) for more information and examples of its application.4. The distributional estimateThe lemma below is adapted from [21, Lemma 2.1].Lemma 6. We have the following bound, 1 2 (log log γ O((log ) )log E ζ(α, X) 2 C(α) 1/α (log ) 1 C( 1 ) 2 log(α 1 ) 1/222for α 1for 12 α 1 .for α 1/2Proof. Note thatE[ ζ(α, X) 2 Yph X(p) 2 iE (p) with E (p) E 1 α.pFor p (2 )1/α we have the trivial bound E (p) (1 1/pα ) 2 . For p (2 )1/α wenotice thatZ π eiθ e iθ 11 α· 1 αdθ I0 (2 /pα )(1 O( /p2α ))E (p) 2π πppPwhere I0 (z) is the 0-th modified Bessel function, and I0 (t) n (t/2)2n /(n!)2 . Inparticular log I0 (t) t2 for 0 t 1. Combining these bounds we get XX 21 2 log E[ ζ(α, X) ] 2 log 1 α C.pp2α1/α1/αp (2 )p (2 )When α 1 the first sum contributes 2 (log log γ O(1/ log )) while the secondcontributes O(1). When 1/2 α 1 the prime number theorem shows that theabove sum is 1/α αα · .log 1 α 2α 1Finally when α tends to 1/2 we use the more careful boundX 2 2 · log(α 21 ) 12αppwhich comes from ζ(2α) 1/(2α 1) O(1), to conclude.

8MARK LEWKO AND MAKSYM RADZIWILL5. Proof of Theorem 2Let N {n1 , . . . , nk }. LetD(X) : Xc(n)X(n).n NConsider the expressionhi22E ζ(α, X) · D(X) .On the one hand, expanding the square this is equal to XXn 1(9)k nk Nc(k) ·kαnα2 Xm,n N ζ(2α)On the other hand, for any , V 0, and(10)c(m)c(n) ·12(km)α· ζ(2α)[k, m]2αX (ni , nj )2α· c(ni )c(nj ).(ni nj )αi,j k α 1, ζ(α, X)D(X) 2 e2V · D(X) 2 k · ζ(α, X) 2( 1) · e 2 V .Indeed to prove this inequality note that if ζ(α, X) eV then the left-hand side isless than e2V · D(X) 2 , while if ζ(α, X) eV , then the left-hand side is less than D(X) 2 · ζ(α, X) 2( 1) · e 2 V k · ζ(α, X) 2( 1) · e 2 V ,using the L bound, D(X) 2 k · kck k coming from Cauchy-Schwarz. Takingthe expectation on both sides and using (9) we get(11)ζ(2α) ·1 X (ni , nj )2α e2V · kck22 k · E[ ζ(α, X) 2 2 ] · e 2 V .k i,j k (ni nj )αIn the above equation, if α 1 then we letV log log log k γ 2/ψ(k) and ψ(k) · log kwith ψ(k) very slowly as k (say ψ(k) log log log k). Otherwise, we let((C(α) log k)1 α · (log log k) α if 21 α 1log kV with .1111/2Vif α 2 log k(C( 2 ) log k log log k) ,With this choice of parameters we have E[ ζ(α, X) 2 2 ] · e 2 V k 1 for a fixed1/2 α 1 and for α 1/2 1/ log k. If 1/2 α 1 is fixed, then inserting the

GCD SUMS AND APPLICATIONS9choice of and V made above into (11) gives the claim. In order to prove the claimfor α 12 , we use Holder’s inequality,1 X (ni , nj ) 1 X (ni , nj )2α 1/(2α) 1 1/(2α) ·k k i,j k ni njk i,j k (ni nj )αwith α 1/2 1/ log k, and appeal to (11) with the choice of parameters as describedabove. The result for 0 α 21 follows in the same manner by interpolating withthe case α 21 using Holder’s inequality.6. Proof of Corollary 1We have already established the upper bound in Theorem 1. Therefore it suffices toobtain the lower bound. Let P(r, ) p 1· . . . · p 1where p1 , p2 , . . . are consecutive1rprimes. Gál proves the following identity in [13], 1XY X (ni , nj )2 v 2 ni njpvv 1ni ,nj P(r, )p p1 .prwhere the summation goes over all ni and nj dividing P(r, )1. The number ofdivisors of P(r, ) is r . Let r be the largest r such that (r log k)r k. Thereforer log k/ log log k and pr log k by the prime number theorem. Pick an integer isuch that(r i)r k (r i 1)r .Since (r log k 1)r 1 k but (r log k)r k it follows that i log k. In particular(r i)r k as k . Set r i and let N {n1 , . . . , nk } be a set containingthe (r i)r k divisors of P(r, ) and k (r i)r other integers picked at random.Then according to Gál’s identity, highlighted above, the GCD sum1 X (ni , nj )2ki,j kni ,nj Nni njis at least, (r i)rY 1 2p pr (1 o(1))kv 1Y p pr1Thisr i 1X1 v ·1 pvr ir i 1X 1 1 2 Yv 1 21 2 1 2· 1 1 ppvr ipp pv 1ridentity can be also quickly checked using the fact that f (m, n) (m, n)2 /(mn) is amultiplicative function of two variables

10MARK LEWKO AND MAKSYM RADZIWILLBy Merten’s theorem the first product is asymptotically equal to (eγ log pr )2 (eγ log log k)2 as k . On the other the second product converges as k to Y X1 1 261 21 2.vppπpv 1Combining these two observations the claim follows.7. Carleson-Hunt bounded variationWe now turn our attention to the proof of Theorem 5. Our argument will dependon the Carleson-Hunt theorem [18], stated below.Proposition 7. There exists an absolute constant c 0 such that!2ZMNXXmaxc(k)e(kx) c(k) 2dx cT1 M Nk 1k 1for any finite sequence (c(k)).We start by writing f as a Fourier seriesXf (x) a(j)e(jx),j Zwhere we have the inequality a(j) (1 j ) 1 . Next we split the Fourier seriesof f into two parts based on the factorization of the j. Here A denotes a large realconstant to be specified later and P (j) corresponds to the largest prime factor of j . LetXXr(x) : a(j)e(jx),p(x) : a(j)e(jx)P (j) (log(N ))2A 2P (j) (log(N ))2A 2so that f (x) p(x) r(x). It suffices to prove, for g(x) {p(x), r(x)} the inequality:!2ZNNXX2(12)maxc(k)g(nk x)dx (log log(N )) c(k) 2 .1 M NTk 1k 1For g(x) p(x) we may write the square root of (12) as 2 1/2ZNXX maxc(k)a(j)e(jnk x) dx T1 M Nk 1 P (j) (log(N ))2A 2

GCD SUMS AND APPLICATIONS ZX a(j) P (j) (log(N ))2A 2maxTNX1 M N11 1/2!2c(k)e(jnk x)dx .k 1Applying the classical Carleson-Hunt inequality this is bounded by!1/2NXX a(j) c(k) 2P (j) (log(N ))2A 2k 1and it remains to notice thatX a(j) P (j) (log(N ))2A 2Xj 1P (j) (log(N ))2A 2 Yp (log(N ))2A 21log log N1 p 1by Merten’s theorem. This completes the analysis of (12).We now consider the left side of (12) with g(x) r(x). The key ingredient in theanalysis of (12) will be the following almost orthogonality property of the functionsr(nk x), which will be proved shortly.Lemma 8. With the notation and conditions stated above, if I [M1 , M2 ] [1, N ]then!2Z X(log log N )2 Xc(k)r(nk x) dx c(k) 2 .2A(logN)Tk Ik IAssuming Lemma 8 for the moment, we may deduce a maximal version of this inequality at the expense of an additional factor of log N using a Radamacher-Menshovtype argument inequality (see [22] for a systematic discussion of this technique). Thislemma will then imply (12) with g(x) r(x) for fixed A 1.Lemma 9. With the notation and conditions stated above,!2ZMNX(log log N )2 X2maxc(k)r(nk x)dx log N · c(k) 2 .2A1 M NlogNTk 1k 1Proof. Without loss of generality assume that N 2n is a power of 2. We considerthe set of diadic subintervals of [1, N ]:D : {[2 (m 1), 2 1 m) : 0 n, 0 m 2n }.Let 1 t(x) N denote the length of the maximal partial sum at x. We maywrite the interval [1, t(x)] as a disjoint union of at most O(log(N )) diadic intervals(elements of D). It follows, for fixed x, that for any t(x) N there exists a disjoint

12MARK LEWKO AND MAKSYM RADZIWILL(x)(x)Ndecomposition of [1, t(x)] into a union of O(log N ) elements {Ds }logs 1 . Here Dsare disjoint dyadic intervals depending on x. Hence 222log Nlog NtXX XX Xc(k)r(nk x) .c(k)r(nk x) c(k)r(nk x) log Ns 1k 1s 1(x)k Dsk DsSumming over all dyadic intervals the dependence on x may be removed. Indeed wehave!2ZZNXX Xc(k)r(nk x)maxlog N c(k)r(nk x) 2T1 M NTk 1log NXZ XD D k Dc(k)r(nk x) 2 dx.T k DD DFinally, Lemma 8 combined with the observation that each integer k occurs inO(log

While estimating (1) for many speci c sequences is straightforward, the problem of determining the maximal size of (1) among all sequences with kterms is much more subtle and the case 1 was posed as a prize problem by the Scienti c Society at Amsterdam in 1947 on Erdos’s suggestion. Date: September 18, 2014.

Related Documents:

2020 pricelist Size/Price in PERENNIALS Page 1 4 Inch 1 Gal 2 Gal 3 Gal 5 gal 7 gal 10 gal 15 Gal 20 gal 30 gal 45 Gal Amsonia 8 12 Big red sage 8 12 Black dalea 8 Black eyed susan 8 Black foot daisy 8 Bluebells 8 Brazos penstemon 8 Butterfly milkweed (A. tuberosa) 8 12 Cardinal flower 8 12 Cedar sage 8 Copper Canyon Daisy 8 12 Cut leaf daisy 8 Daminita 8 12 Dwarf petunia ‘Katie’ 6

ning B 3 tum 20 l (5 gal.) CS Neopren B PROFI-NET C 3 tum 20 l (5 gal.) CM EPDM C PROFI-BUS D 3 tum 20 l (5 gal.) CM Neopren D Device-Net F 3 tum 200 l (55 gal.) CS EPDM N Inget G 3 tum 200 l (55 gal.) CS Neopren H 3 tum 200 l (55 gal.) CM EPDM J 3 tum 200 l (55 gal.) CM Neopren K 6 tum 200 l (55 gal.) CS EPDM M 6 tum 200 l (55 gal.) CS Neopren .

2186 buccaneer 5 extra (265 gal) 265 gal 15.01 3,977.21 2190 buccaneer 5 extra (30 gal) 30 gal 15.30 459.00 2192 buccaneer plus (265) 265 gal 11.28 2,987.88 2198 buccaneer plus (2x2.5) 2.5 gal 12.44 31.10 2195 buccaneer plus (30 gal) 30 gal 11.74 352.25

1 c or 1 gal 3 qt or 3 gal 1 pt or 1 gal 10 qt or 10 gal 5. coffee pot 6. bowl of soup 7. tea kettle 8. small milk carton 1 c or 1 gal 1 pt or 1 gal 1 qt or 1 gal 1 c or 1 gal Choose the better unit to measure the capacity of each. 9. hot tub 10. shampoo bottle 11. bucket 12.

Cylinder stroke 47.6 in (1 210 mm) SERVICE CAPACITIES Fuel tank 153.2 gal (580 L) Hydraulic system 92.5 gal (350 L) Hydraulic tank 46.2 gal (175 L) Engine oil w/ filter 12.7 gal (48 L) Swing drive 1.3 gal (5 L) 2.5 gal (9.5 L) Cooling system 10.0 gal (38 L) DEF tank 40.2 gal (152 L) Page 1 of 7 EXCAVATOR TIER 4 FINAL CERTIFIED

Manual Manual Electric Manual 40W 40W 75W 75W M-240-T-1 * M-240-T1 ' M-240-T-1 ' W-240-T-1 '.020" .020" .020" 020".014"-.018" .014"-.018'" .014"-.018" .014"-.018" 3 gal. 3 gal. 3 gal. 3 gal. 3.75 gal. 3.75gal. 3.75 gal. 3.75 gal. 20:1 20:1 20:1 20:1 BRAKE Type Drum Drum Drum Drum ACCES-SORIES Speedometer Tachometer Optional Optional Optional .

expansion tank sizing chart assumed 140 temp and 150 psi max. pressure expansion tank needed incoming water pressure water heater capacity (gallons) 30 40 50 66 80 100 120 40 psi 2 gal 2 gal 2 gal 5 gal 5 gal 5 gal 5 gal 50

Cyrtomium falcatum Japanese Holly Fern 1 gal Dennstaedtia punctilobula Hay Scented Fern 1 gal Diplazium pycnocarpon Glade Fern 1 gal Dryopteris australis Dixie Wood Fern 1 gal Dryopteris atrata Shaggy Shield Fern 1 gal Dryopteris atrata Shaggy Shield Fern 3 gal Dryopteris erythrosora 'Brilliance' Autumn Brilliance Fern 1 gal