1 Introduction
Our focus in this paper is the Vinogradov system of Diophantine equations. When
$k$
and
$s$
are natural numbers, and
$X$
is a large real number, denote by
$J_{s,k}(X)$
the number of integral solutions of the system

with
$1\leqslant x_{i},y_{i}\leqslant X$
$(1\leqslant i\leqslant s)$
. Bounds for
$J_{s,k}(X)$
are linked via orthogonality to mean value estimates for exponential sums and hence feature prominently in the analysis of problems spanning the analytic theory of numbers. Thus, for example, the strongest available conclusions concerning the validity of the asymptotic formula in Waring’s problem, the zero-free region for the Riemann zeta function and the distribution modulo
$1$
of polynomial sequences all depend on bounds for
$J_{s,k}(X)$
. We refer the reader to the comprehensive sources due to Arkhipov et al [Reference Arkhipov, Chubarikov and Karatsuba1] and Vinogradov [Reference Vinogradov11] for discussion of the many applications of such estimates, collectively described as instances of Vinogradov’s mean value theorem. A more recent tour of the subject is described in [Reference Wooley18]. By considering mean values of an associated exponential sum, it follows that the validity of the strongly diagonal estimate

for
$1\leqslant s\leqslant \frac{1}{2}k(k+1)$
, would imply the main conjecture in Vinogradov’s mean value theorem, which asserts that for each
$\unicode[STIX]{x1D700}>0$
, one has

Here and throughout this paper, the implicit constants associated with Vinogradov’s notation
$\ll$
and
$\gg$
may depend on
$s$
,
$k$
and
$\unicode[STIX]{x1D700}$
.
For almost all of the eighty year history of the subject, an estimate of the strength of (1.3) has seemed a very remote prospect indeed, for until only a year before this paper was written, the bound (1.2) was known to hold only for
$s\leqslant k+1$
(see [Reference Hua8, Lemma 5.4]). By enhancing the efficient congruencing method introduced in [Reference Wooley15], recent work of the author joint with Ford [Reference Ford and Wooley4] has established the estimate (1.2) for
$s\leqslant \frac{1}{4}(k+1)^{2}$
, amounting to half of the range
$s\leqslant \frac{1}{2}k(k+1)$
in which diagonal behaviour dominates (1.3). Our goal in this paper is to adapt the multigrade efficient congruencing method detailed in our very recent work [Reference Wooley19] so as to obtain strongly diagonal estimates of the shape (1.2). It transpires that when
$k$
is large, we are able to establish the validity of the conjectured estimate (1.2) for almost all
$s$
in the critical interval
$1\leqslant s\leqslant \frac{1}{2}k(k+1)$
or, as one might say, a proportion asymptotically approaching
$100\%$
of the critical interval. We consequently find ourselves within a whisker of the proof of the main conjecture (1.3), a situation hitherto unprecedented in the analysis of mean value estimates for exponential sums of large degree.
The most precise form of our new estimate may be found in Theorem 2.1 below. For the present, we restrict ourselves to a conclusion relatively simple to state that is indicative of what may now be achieved.
Theorem 1.1. Let
$k$
be a natural number with
$k\geqslant 7$
, put
$r=k-\lceil 2\sqrt{k}\rceil +2$
and suppose that
$s$
is an integer satisfying

Then, for each
$\unicode[STIX]{x1D700}>0$
, one has
$J_{s,k}(X)\ll X^{s+\unicode[STIX]{x1D700}}$
.
A modicum of computation reveals that when
$k$
is large, this estimate confirms the main conjecture (1.3) for almost all of the critical interval of exponents
$1\leqslant s\leqslant {\textstyle \frac{1}{2}}k(k+1)$
.
Corollary 1.2. When
$s$
and
$k$
are natural numbers with
$k\geqslant 7$
and

then, for each
$\unicode[STIX]{x1D700}>0$
, one has
$J_{s,k}(X)\ll X^{s+\unicode[STIX]{x1D700}}$
.
As we have already noted, a conclusion analogous to Theorem 1.1 is made available in [Reference Ford and Wooley4, Theorem 1.1] in the more limited range
$1\leqslant s\leqslant \frac{1}{4}(k+1)^{2}$
. Earlier conclusions were limited to the much shorter interval
$1\leqslant s\leqslant k+1$
(see [Reference Vaughan and Wooley10] for a particularly precise statement in this situation). In view of the lower bound

that arises by considering the diagonal solutions of (1.1) with
$\mathbf{x}=\mathbf{y}$
, together with a lower bound for the product of local densities (see [Reference Vaughan9, equation (7.4)]), one sees that the conclusion of Theorem 1.1 cannot hold when
$s>{\textstyle \frac{1}{2}}k(k+1)$
.
The gap between our new result and a complete proof of the main conjecture (1.3) stands at
${\textstyle \frac{7}{3}}k$
variables, amounting to a proportion
$O(1/k)$
of the length of the critical interval
$1\leqslant s\leqslant \frac{1}{2}k(k+1)$
. In a certain sense, therefore, our conclusion establishes the main conjecture in 100% of the critical interval as
$k\rightarrow \infty$
, justifying the assertion concluding our opening paragraph. In §11, we describe some modifications to our basic strategy which offer some improvement in the conclusion of Corollary 1.2 towards the main conjecture (1.3). We have opted for a relatively concise account of this improvement, since the details of the associated argument are of sufficient complexity that the key elements of our basic strategy would be obscured were we to make this account the main focus of our exposition.
Theorem 1.3. When
$k$
is a sufficiently large natural number and
$s$
is an integer satisfying

then, for each
$\unicode[STIX]{x1D700}>0$
, one has
$J_{s,k}(X)\ll X^{s+\unicode[STIX]{x1D700}}$
.
Our methods are applicable not only for large values of
$k$
. By taking
$r=k-\lceil 2\sqrt{k}\rceil +2$
in Theorem 2.1 below, one may establish an improvement on the conclusion of [Reference Ford and Wooley4, Theorem 1.1]. We note here that the latter delivers the bound
$J_{s,k}(X)\ll X^{s+\unicode[STIX]{x1D700}}$
for
$1\leqslant s\leqslant \frac{1}{4}(k+1)^{2}$
.
Theorem 1.4. Define the number
$D(k)$
as in Table 1. Then, whenever
$k\geqslant 4$
and
$1\leqslant s\leqslant D(k)$
, one has
$J_{s,k}(X)\ll X^{s+\unicode[STIX]{x1D700}}$
.
Table 1 Values for
$D(k)$
described in Theorem 1.4.

One may enquire concerning the impact of our methods in the case
$k=3$
. Here, although the methods of this paper do not apply directly, a careful variant of our ideas proves surprisingly successful. This is a matter which we address elsewhere (see [Reference Wooley20]).
A trivial consequence of the estimate supplied by Theorems 1.1 and 1.3 provides a new upper bound for
$J_{s,k}(X)$
at the critical value
$s=\frac{1}{2}k(k+1)$
.
Theorem 1.5. Suppose that
$k$
is a large natural number. Then

where
$\unicode[STIX]{x1D6E5}=({\textstyle \frac{1}{3}}+o(1))k$
.
As remarked in the discussion following the statement of [Reference Ford and Wooley4, Theorem 1.2], the work of that paper shows that a conclusion analogous to Theorem 1.5 holds with
$\unicode[STIX]{x1D6E5}=(\frac{3}{2}-\sqrt{2})k^{2}+O(k)$
. Our new result for the first time in the subject obtains a conclusion in which
$\unicode[STIX]{x1D6E5}=o(k^{2})$
. Indeed, one now has a good asymptotic approximation to the main conjecture (1.3) for all
$s$
, since we have

with
$\unicode[STIX]{x1D6E5}_{s,k}=O(k)$
for all
$s$
.
We are also able to obtain a very slight improvement in our previous bound for the the number of variables required to establish the anticipated asymptotic formula in Waring’s problem, at least, when
$k$
is sufficiently large. We refer the reader to Theorem 12.2 for an account of this new bound, and to Theorem 12.1 for some consequences of Theorem 1.3 in the context of Tarry’s problem. The latter is worth highlighting already. When
$h$
,
$k$
and
$s$
are positive integers with
$h\geqslant 2$
, consider the Diophantine system

Let
$W(k,h)$
denote the least natural number
$s$
having the property that the simultaneous equations (1.5) possess an integral solution
$\mathbf{x}$
with

Also, let
$P(k,h)$
denote the least natural number
$s$
having the property that the simultaneous equations (1.5) possess an integral solution
$\mathbf{x}$
with

It is trivial that
$P(k,h)\leqslant W(k,h)$
, and relatively straightforward to establish that when
$h$
and
$k$
are natural numbers with
$h\geqslant 2$
, then
$P(k,h)\leqslant {\textstyle \frac{1}{2}}k(k+1)+1$
(see Hardy and Wright [Reference Hardy and Wright5, Theorem 409]). The corresponding bound for the more difficult quantity
$W(k,h)$
has, however, hitherto remained far beyond reach. By applying Theorem 1.5, we are now able to show in Theorem 12.1 that when
$h$
and
$k$
are natural numbers with
$h\geqslant 2$
and
$k$
sufficiently large, then one has
$W(k,h)\leqslant {\textstyle \frac{1}{2}}k(k+1)+1$
. As noted earlier, other applications of Vinogradov’s mean value theorem can be found in the classical literature on the subject, such as [Reference Arkhipov, Chubarikov and Karatsuba1, Reference Vinogradov11, Reference Wooley12].
We have written elsewhere concerning the basic strategy underlying efficient congruencing; a sketch of the basic ideas can be found in [Reference Wooley15, §2]. The multigrade efficient congruencing method seeks to more fully extract the information available from congruences of different levels. We direct the reader to [Reference Wooley19] for an introduction to such ideas. Perhaps it is worth noting that the main challenge in the present adaptation of multigrade efficient congruencing is that no loss of more than a constant factor can be tolerated in the basic congruencing steps. This task is complicated as we progress through the iteration, stepping from congruences modulo
$p^{a}$
to congruences modulo
$p^{b}$
, since the ratio
$b/a$
can vary widely according to which of the several possible moduli dominate proceedings at each stage of the iteration. Finally, we remark that although one can envision further improvement in the range of
$s$
accommodated by Theorem 1.3, it would seem that a new idea is required to replace the defect
${\textstyle \frac{1}{3}}k$
by a quantity
$ck$
with
$c<{\textstyle \frac{1}{3}}$
. The term
${\textstyle \frac{1}{3}}k$
seems to be an unavoidable consequence of the application of Hölder’s inequality underlying Lemmas 7.1 and 11.3 that teases apart the information available from congruences of different levels.
Recent developments (added 29th May 2016)
The area of research focusing on Vinogradov’s mean value theorem and its hinterland has recently undergone a burst of activity, with rapid progress from several quarters. On the one hand, the author [Reference Wooley20] has applied the methods of this paper to establish the cubic case of the main conjecture in full. On the other hand, very recent work of Bourgain et al [Reference Bourgain, Demeter and Guth2] extended the range

in which our Theorem 1.3 applies, to the full critical range
$1\leqslant s\leqslant {\textstyle \frac{1}{2}}k(k+1)$
, thereby confirming the main conjecture in full. The latter work [Reference Bourgain, Demeter and Guth2] was submitted to the arXiv on 14th December 2015, the most recent (third) revision having been submitted on 1st April 2016. This paper is in all essential mathematical respects identical with a version predating the work of Bourgain et al, and was submitted to the arXiv on 13th January 2014. However, its appearance in print has been delayed by forces beyond the author’s control.
The methods of this paper may be seen as a
$p$
-adic version of the method underlying [Reference Bourgain, Demeter and Guth2], the latter being more akin to Vinogradov’s use of real short intervals dating from the 1930s. Two comments are in order here. First, the
$p$
-adic methods of the present work continue to represent a useful alternative to those of [Reference Bourgain, Demeter and Guth2], since they are easily adapted to other number-theoretic situations, such as those relevant to number fields and function fields. Second, it seems now very likely that the methods of the present paper, in combination with subsequent developments, will also establish the main conjecture in full, in all the settings just outlined as well as in the real setting explored in [Reference Bourgain, Demeter and Guth2]. It seems reasonable to speculate that methods of the widest applicability will incorporate elements of both approaches.
2 The basic infrastructure
We begin our discussion of the proof of Theorem 1.1 by describing the components and basic notation that we subsequently assemble into the apparatus required for the multigrade efficient congruencing method. This is very similar to that introduced in our recent work [Reference Wooley19], and resembles the infrastructure of our earliest work on efficient congruencing [Reference Wooley15]. The reader might wonder whether some kind of universal account could be given that would encompass all possibilities. For the present such a goal seems premature, since innovations are the rule for such a rapidly evolving circle of ideas. Instead, novel consequences of the method seem to require careful arrangement of parameters together with a significant measure of artistry.
We establish Theorem 1.1 as a consequence of a theorem that in many circumstances is somewhat sharper, though more complicated to state.
Theorem 2.1. Suppose that
$s$
,
$k$
and
$r$
are natural numbers with
$k\geqslant 4$
,

where

and

Then, for each
$\unicode[STIX]{x1D700}>0$
, one has
$J_{s+r}(X)\ll X^{s+r+\unicode[STIX]{x1D700}}$
.
We consider
$k$
to be fixed, and abbreviate
$J_{s,k}(X)$
to
$J_{s}(X)$
without further comment. Let
$s\in \mathbb{N}$
be arbitrary, and define
$\unicode[STIX]{x1D706}_{s}^{\ast }\in \mathbb{R}$
by means of the relation

Thus, for each
$\unicode[STIX]{x1D700}>0$
, and any
$X\in \mathbb{R}$
sufficiently large in terms of
$s$
,
$k$
and
$\unicode[STIX]{x1D700}$
, one has
$J_{s}(X)\ll X^{\unicode[STIX]{x1D706}_{s}^{\ast }+\unicode[STIX]{x1D700}}$
. Note that the lower bound (1.4) combines with a trivial estimate for
$J_{s}(X)$
to show that
$s\leqslant \unicode[STIX]{x1D706}_{s}^{\ast }\leqslant 2s$
, whilst the conjectured upper bound (1.2) implies that
$\unicode[STIX]{x1D706}_{s}^{\ast }=s$
for
$s\leqslant \frac{1}{2}k(k+1)$
.
Next, we recall some standard notational conventions. The letters
$s$
and
$k$
denote natural numbers with
$k\geqslant 3$
, and
$\unicode[STIX]{x1D700}$
denotes a sufficiently small positive number. Our basic parameter is
$X$
, a large real number depending at most on
$k$
,
$s$
and
$\unicode[STIX]{x1D700}$
, unless otherwise indicated. Whenever
$\unicode[STIX]{x1D700}$
appears in a statement, we assert that the statement holds for each
$\unicode[STIX]{x1D700}>0$
. As usual, we write
$\lfloor \unicode[STIX]{x1D713}\rfloor$
to denote the largest integer no larger than
$\unicode[STIX]{x1D713}$
, and
$\lceil \unicode[STIX]{x1D713}\rceil$
to denote the least integer no smaller than
$\unicode[STIX]{x1D713}$
. We make sweeping use of vector notation. Thus, with
$t$
implied from the environment at hand, we write
$\mathbf{z}\equiv \mathbf{w}~(\text{mod}~p)$
to denote that
$z_{i}\equiv w_{i}~(\text{mod}~p)$
$(1\leqslant i\leqslant t)$
, or
$\mathbf{z}\equiv \unicode[STIX]{x1D709}~(\text{mod}~p)$
to denote that
$z_{i}\equiv \unicode[STIX]{x1D709}~(\text{mod}~p)$
$(1\leqslant i\leqslant t)$
, or
$[\mathbf{z}~(\text{mod}~q)]$
to denote the
$t$
-tuple
$(\unicode[STIX]{x1D701}_{1},\ldots ,\unicode[STIX]{x1D701}_{t})$
, where for
$1\leqslant i\leqslant t$
one has
$1\leqslant \unicode[STIX]{x1D701}_{i}\leqslant q$
and
$z_{i}\equiv \unicode[STIX]{x1D701}_{i}~(\text{mod}~q)$
. Finally, we employ the convention that whenever
$G: [0,1)^{k}\rightarrow \mathbb{C}$
is integrable, then

Thus, on writing

where as usual
$e(z)$
denotes
$\text{e}^{2\unicode[STIX]{x1D70B}\text{i}z}$
, it follows from orthogonality that

We next introduce the parameters which define the iterative method that is the central concern of this paper. We suppose throughout that
$k\geqslant 4$
, and put

We take
$r$
to be an integral parameter satisfying
$1\leqslant r\leqslant r_{0}$
, and we consider a natural number
$s$
with
$k+1\leqslant s\leqslant {\textstyle \frac{1}{2}}k(k+1)$
. We then define
$\unicode[STIX]{x1D6E5}=\unicode[STIX]{x1D6E5}(r,s)$
by taking

and suppose in what follows that

Next, we put

and then define

We will suppose throughout that
$s<\unicode[STIX]{x1D703}_{+}$
. Here, some explanation may defuse confusion that may arise from the implicit dependence of
$\unicode[STIX]{x1D703}_{+}$
on
$s$
. In practice, the definition (2.3) ensures that
$\unicode[STIX]{x1D6E5}$
is no larger than about
${\textstyle \frac{1}{3}}k$
, and hence one finds from (2.6) that
$\unicode[STIX]{x1D703}_{+}$
is roughly

The condition
$s<\unicode[STIX]{x1D703}_{+}$
is consequently easily verified, since there are relatively few values of
$s$
to check, and in particular none exceeding
${\textstyle \frac{1}{2}}k(k+1)$
. Likewise, the condition (2.4) is also easily verified for the values of
$s$
open to discussion.
Our goal is to establish that
$\unicode[STIX]{x1D706}_{s+r}^{\ast }=s+r$
. Having established the latter, one finds by applying Hölder’s inequality to the right-hand side of (2.1) that whenever
$1\leqslant t\leqslant s+r$
, one has

whence
$\unicode[STIX]{x1D706}_{t}^{\ast }=t$
. Thus, it is clear that the main conclusions of this paper follow by focusing on
$\unicode[STIX]{x1D706}_{s+r}^{\ast }$
. Henceforth, for brevity we write
$\unicode[STIX]{x1D706}=\unicode[STIX]{x1D706}_{s+r}^{\ast }$
.
Let
$R$
be a natural number sufficiently large in terms of
$s$
and
$k$
. Specifically, we choose
$R$
as follows. From (2.5) and (2.6), one finds that
$\unicode[STIX]{x1D703}_{+}\unicode[STIX]{x1D703}_{-}=\mathfrak{b}$
, and so the hypothesis
$s<\unicode[STIX]{x1D703}_{+}$
ensures that

Observe next that
$\unicode[STIX]{x1D703}_{+}>\unicode[STIX]{x1D703}_{-}$
, so that on writing
$\unicode[STIX]{x1D70F}=\unicode[STIX]{x1D703}_{+}/\unicode[STIX]{x1D703}_{-}$
, one finds that

Hence, we obtain the upper bound

Recall again our assumption that
$s<\unicode[STIX]{x1D703}_{+}$
, and put
$\unicode[STIX]{x1D708}=\unicode[STIX]{x1D703}_{+}-s>0$
. Then we have

Consequently, if we take
$R=\lceil W\unicode[STIX]{x1D703}_{+}/\unicode[STIX]{x1D708}\rceil$
, with
$W$
a large enough integer, then we ensure that

whence

The significance of this condition will become apparent in due course (see equation (8.1) below, and the ensuing discussion).
Having fixed
$R$
satisfying these conditions, we take
$N$
to be a natural number sufficiently large in terms of
$s$
,
$k$
and
$R$
, and then put

In view of the definition of
$\unicode[STIX]{x1D706}$
, there exists a sequence of natural numbers
$(X_{l})_{l=1}^{\infty }$
, tending to infinity, with the property that
$J_{s+r}(X_{l})>X_{l}^{\unicode[STIX]{x1D706}-\unicode[STIX]{x1D6FF}}$
$(l\in \mathbb{N})$
. Also, provided that
$X_{l}$
is sufficiently large, one has the corresponding upper bound
$J_{s+r}(Y)<Y^{\unicode[STIX]{x1D706}+\unicode[STIX]{x1D6FF}}$
for
$Y\geqslant X_{l}^{1/2}$
. We consider a fixed element
$X=X_{l}$
of the sequence
$(X_{l})_{l=1}^{\infty }$
, which we may assume to be sufficiently large in terms of
$s$
,
$k$
and
$N$
. We put
$M=X^{\unicode[STIX]{x1D703}}$
, and note from (2.8) that
$X^{\unicode[STIX]{x1D6FF}}<M^{1/N}$
. Throughout, implicit constants may depend on
$s$
,
$k$
,
$N$
, and also on
$\unicode[STIX]{x1D700}$
in view of our earlier convention, but not on any other variable.
We next introduce the cast of exponential sums and mean values appearing in our arguments. Let
$p$
be a prime number with
$M<p\leqslant 2M$
to be fixed in due course. When
$c$
and
$\unicode[STIX]{x1D709}$
are non-negative integers, and
$\unicode[STIX]{x1D736}\in [0,1)^{k}$
, we define

When
$1\leqslant m\leqslant k-1$
, denote by
$\unicode[STIX]{x1D6EF}_{c}^{m}(\unicode[STIX]{x1D709})$
the set of integral
$m$
-tuples
$(\unicode[STIX]{x1D709}_{1},\ldots ,\unicode[STIX]{x1D709}_{m})$
, with
$1\leqslant \unicode[STIX]{x1D743}\leqslant p^{c+1}$
and
$\unicode[STIX]{x1D743}\equiv \unicode[STIX]{x1D709}~(\text{mod}~p^{c})$
, and satisfying the property that
$\unicode[STIX]{x1D709}_{i}\not \equiv \unicode[STIX]{x1D709}_{j}~(\text{mod}~p^{c+1})$
for
$i\neq j$
. We then put

When
$a$
and
$b$
are positive integers, we define


This notation extends to the case
$m=0$
, of course, by simply interpreting the exponential sum
$\mathfrak{F}_{c}^{0}(\unicode[STIX]{x1D736};\unicode[STIX]{x1D709})$
to be
$1$
.
Note that, by orthogonality, the mean value
$I_{a,b}^{m,r}(X;\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
counts the number of integral solutions of the system

with
$1\leqslant \mathbf{x},\mathbf{y},\mathbf{v},\mathbf{w}\leqslant X$
,
$\mathbf{v}\equiv \mathbf{w}\equiv \unicode[STIX]{x1D702}~(\text{mod}~p^{b})$
and
$\mathbf{x},\mathbf{y}\in \unicode[STIX]{x1D6EF}_{a}^{m}(\unicode[STIX]{x1D709})~(\text{mod}~p^{a+1})$
. Similarly, the mean value
$K_{a,b}^{m,r}(X;\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
counts the number of solutions of

with
$1\leqslant \mathbf{x},\mathbf{y},\mathbf{u},\mathbf{v},\mathbf{w},\mathbf{z}\leqslant X$
, and satisfying
$\mathbf{w}\equiv \mathbf{z}\equiv \unicode[STIX]{x1D702}~(\text{mod}~p^{b})$
,

Given any one such solution of the system (2.14), an application of the binomial theorem shows that
$\mathbf{x}-\unicode[STIX]{x1D702}$
,
$\mathbf{y}-\unicode[STIX]{x1D702}$
,
$\mathbf{u}-\unicode[STIX]{x1D702}$
,
$\mathbf{v}-\unicode[STIX]{x1D702}$
,
$\mathbf{w}-\unicode[STIX]{x1D702}$
,
$\mathbf{z}-\unicode[STIX]{x1D702}$
is also a solution. Since in any solution counted by
$K_{a,b}^{m,r}(X;\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
, one has
$\mathbf{u}\equiv \mathbf{v}\equiv \unicode[STIX]{x1D702}~(\text{mod}~p^{b})$
and
$\mathbf{w}\equiv \mathbf{z}\equiv \unicode[STIX]{x1D702}~(\text{mod}~p^{b})$
, we deduce that

We put


The implicit dependence on
$p$
in the above notation will be rendered irrelevant in §5, since we fix the choice of this prime following the proof of Lemma 5.1.
We must align the definition of
$K_{0,b}^{m,r}(X)$
with the conditioning idea. When
$\boldsymbol{\unicode[STIX]{x1D701}}$
is a tuple of integers, we denote by
$\unicode[STIX]{x1D6EF}^{m}(\boldsymbol{\unicode[STIX]{x1D701}})$
the set of
$m$
-tuples
$(\unicode[STIX]{x1D709}_{1},\ldots ,\unicode[STIX]{x1D709}_{m})\in \unicode[STIX]{x1D6EF}_{0}^{m}(0)$
such that
$\unicode[STIX]{x1D709}_{i}\not \equiv \unicode[STIX]{x1D701}_{j}~(\text{mod}~p)$
for all
$i$
and
$j$
. Recalling (2.9), we put

and then define



As in our earlier work, we make use of an operator that makes transparent the relationship between mean values and their anticipated magnitudes. For the purposes at hand, it is useful to reconfigure this normalization so as to make visible the deviation in a mean value from strongly diagonal behaviour. Thus, in the present circumstances, we adopt the convention that



Using this notation, our earlier bounds for
$J_{s+r}(X)$
may be rewritten in the form

where
$\unicode[STIX]{x1D6EC}$
is defined by
$\unicode[STIX]{x1D6EC}=\unicode[STIX]{x1D706}-(s+r)$
.
Finally, we recall two simple estimates associated with the system (1.1).
Lemma 2.2. Suppose that
$c$
is a non-negative integer with
$c\unicode[STIX]{x1D703}\leqslant 1$
. Then, for each natural number
$u$
, one has

Proof. This is [Reference Wooley15, Lemma 3.1]. ◻
Lemma 2.3. Suppose that
$c$
and
$d$
are non-negative integers with
$c\leqslant \unicode[STIX]{x1D703}^{-1}$
and
$d\leqslant \unicode[STIX]{x1D703}^{-1}$
. Then, whenever
$u,v\in \mathbb{N}$
and
$\unicode[STIX]{x1D709},\unicode[STIX]{x1D701}\in \mathbb{Z}$
, one has

Proof. This is [Reference Ford and Wooley4, Corollary 2.2]. ◻
3 Auxiliary systems of congruences
Our goal of establishing strongly diagonal behaviour dictates that we must establish essentially diagonal behaviour for the solutions of auxiliary congruences central to our methods. Complicating our discussion in present circumstances is the potential for the ratio
$b/a$
to be small, for in earlier work [Reference Ford and Wooley4, Reference Wooley17] this was assumed to be at least
$\frac{1}{2}(k-1)$
or thereabouts. We begin by adjusting our notation to accommodate the demands of this paper. When
$a$
and
$b$
are integers with
$1\leqslant a<b$
, we write
${\mathcal{B}}_{a,b}^{n}(\mathbf{m};\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
for the set of solutions of the system of congruences

with
$1\leqslant \mathbf{z}\leqslant p^{kb}$
and
$\mathbf{z}\equiv \unicode[STIX]{x1D743}~(\text{mod}~p^{a+1})$
for some
$\unicode[STIX]{x1D743}\in \unicode[STIX]{x1D6EF}_{a}^{n}(\unicode[STIX]{x1D709})$
. We define an equivalence relation
${\mathcal{R}}(\unicode[STIX]{x1D706})$
on integral
$n$
-tuples by declaring
$\mathbf{x}$
and
$\mathbf{y}$
to be
${\mathcal{R}}(\unicode[STIX]{x1D706})$
-equivalent when
$\mathbf{x}\equiv \mathbf{y}~(\text{mod}~p^{\unicode[STIX]{x1D706}})$
. When
$hb\in \mathbb{N}$
, we then write
${\mathcal{C}}_{a,b}^{n,h}(\mathbf{m};\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
for the set of
${\mathcal{R}}(hb)$
-equivalence classes of
${\mathcal{B}}_{a,b}^{n}(\mathbf{m};\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
, and we define
$B_{a,b}^{n,h}(p)$
by putting

When
$a=0$
we modify these definitions, so that
${\mathcal{B}}_{0,b}^{n}(\mathbf{m};\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
denotes the set of solutions of the system of congruences (3.1) with
$1\leqslant \mathbf{z}\leqslant p^{kb}$
and
$\mathbf{z}\equiv \unicode[STIX]{x1D743}~(\text{mod}~p)$
for some
$\unicode[STIX]{x1D743}\in \unicode[STIX]{x1D6EF}_{0}^{n}(\unicode[STIX]{x1D709})$
, and for which in addition
$\mathbf{z}\not \equiv \unicode[STIX]{x1D702}~(\text{mod}~p)$
. As in the situation in which one has
$a\geqslant 1$
, when
$hb\in \mathbb{N}$
, we write
${\mathcal{C}}_{0,b}^{n,h}(\mathbf{m};\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
for the set of
${\mathcal{R}}(hb)$
-equivalence classes of
${\mathcal{B}}_{0,b}^{n}(\mathbf{m};\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
, but we define
$B_{0,b}^{n,h}(p)$
by putting

We observe that the definition of
$B_{a,b}^{n,h}(p)$
ensures that whenever
$0<h^{\prime }\leqslant h$
and both
$hb$
and
$h^{\prime }b$
are natural numbers, one has

Readers might wonder concerning the wisdom of permitting in
$B_{a,b}^{n,h}(p)$
the parameter
$h$
to be a rational number with
$hb\in \mathbb{N}$
. We remark that this somewhat strange notational device is designed to facilitate ease of comparison with the earlier work [Reference Ford and Wooley4, Reference Wooley17, Reference Wooley19]. Since the role of the parameter
$h$
is simply that of indicating certain congruence relations modulo
$p^{hb}$
, the reader may take comfort from our imposition of the condition
$hb\in \mathbb{N}$
.
Lemma 3.1. Let
$m$
be an integer with
$0\leqslant m\leqslant k-1$
, and suppose that
$a$
and
$b$
are integers with
$0\leqslant a<b$
and
$(k-m)b\geqslant (m+1)a$
. Suppose in addition that
$h$
is a natural number with
$a\leqslant h\leqslant (k-m)b-ma$
. Then one has

Proof. We are able to follow the arguments of the proofs of [Reference Wooley17, Lemma 3.5] and [Reference Ford and Wooley4, Lemma 3.3], with some modifications. For the sake of transparency of exposition, we provide an essentially complete account. We begin by considering the situation with
$a\geqslant 1$
, the alterations required when
$a=0$
being easily accommodated within our basic argument. Consider fixed natural numbers
$a$
and
$b$
with
$1\leqslant a<b$
and
$(k-m)b\geqslant (m+1)a$
, and fixed integers
$\unicode[STIX]{x1D709}$
and
$\unicode[STIX]{x1D702}$
with
$1\leqslant \unicode[STIX]{x1D709}\leqslant p^{a}$
,
$1\leqslant \unicode[STIX]{x1D702}\leqslant p^{b}$
and
$\unicode[STIX]{x1D702}\not \equiv \unicode[STIX]{x1D709}~(\text{mod}~p)$
. Note that in view of the relation (3.4), the conclusion of the lemma follows in general if we can establish it when
$h=(k-m)b-ma$
. We henceforth assume that the latter is the case. We then denote by
${\mathcal{D}}_{1}(\mathbf{n})$
the set of
${\mathcal{R}}(h)$
-equivalence classes of solutions of the system of congruences

with
$1\leqslant \mathbf{z}\leqslant p^{kb}$
and
$\mathbf{z}\equiv \unicode[STIX]{x1D743}~(\text{mod}~p^{a+1})$
for some
$\unicode[STIX]{x1D743}\in \unicode[STIX]{x1D6EF}_{a}^{m+1}(\unicode[STIX]{x1D709})$
. To any solution
$\mathbf{z}$
of (3.1) there corresponds a unique
$(m+1)$
-tuple
$(n_{k-m},\ldots ,n_{k})$
, satisfying
$1\leqslant \mathbf{n}\leqslant p^{(k-m)b}$
, for which (3.5) holds and

Since
$h/b\leqslant k-m$
, we therefore infer that

We fix any choice of
$\mathbf{n}$
for which the maximum is achieved in (3.6). There is plainly no loss of generality in supposing that
${\mathcal{D}}_{1}(\mathbf{n})$
is non-empty. Observe that for any solution
$\mathbf{z}^{\prime }$
of (3.5) there is an
${\mathcal{R}}(k-m)$
-equivalent solution
$\mathbf{z}$
satisfying
$1\leqslant \mathbf{z}\leqslant p^{(k-m)b}$
. Rewriting each variable
$z_{i}$
in (3.5) in the shape
$z_{i}=p^{a}y_{i}+\unicode[STIX]{x1D709}$
, we infer from the hypothesis that
$\mathbf{z}\equiv \unicode[STIX]{x1D743}~(\text{mod}~p^{a+1})$
for some
$\unicode[STIX]{x1D743}\in \unicode[STIX]{x1D6EF}_{a}^{m+1}(\unicode[STIX]{x1D709})$
that the
$(m+1)$
-tuple
$\mathbf{y}$
necessarily satisfies

Write
$\unicode[STIX]{x1D701}=\unicode[STIX]{x1D709}-\unicode[STIX]{x1D702}$
, and note that the constraint
$\unicode[STIX]{x1D702}\not \equiv \unicode[STIX]{x1D709}~(\text{mod}~p)$
ensures that
$p\nmid \unicode[STIX]{x1D701}$
. We denote the multiplicative inverse of
$\unicode[STIX]{x1D701}$
modulo
$p^{(k-m)b}$
by
$\unicode[STIX]{x1D701}^{-1}$
. Then we deduce from (3.5) that
$\text{card}({\mathcal{D}}_{1}(\mathbf{n}))$
is bounded above by the number of
${\mathcal{R}}(h-a)$
-equivalence classes of solutions of the system of congruences

with
$1\leqslant \mathbf{y}\leqslant p^{(k-m)b-a}$
satisfying (3.7). Let
$\mathbf{y}=\mathbf{w}$
be any solution of the system (3.8). Then we find that all other solutions
$\mathbf{y}$
satisfy the system

We next apply [Reference Wooley17, Lemma 3.2], just as in the arguments of the proofs of [Reference Wooley17, Lemmas 3.3 to 3.6]. Consider an index
$j$
with
$k-m\leqslant j\leqslant k$
, and apply the latter lemma with
$\unicode[STIX]{x1D6FC}=k-m-1$
and
$\unicode[STIX]{x1D6FD}=j-k+m+1$
. We find that there exist integers
$c_{j,l}$
$(k-m-1\leqslant l\leqslant j)$
and
$d_{j,u}$
$(j-k+m+1\leqslant u\leqslant j)$
, bounded in terms of
$k$
, and with
$d_{j,j-k+m+1}\neq 0$
, for which one has the polynomial identity

Since we may assume
$p$
to be large, we may suppose that
$p\nmid d_{j,j-k+m+1}$
. Thus, multiplying (3.10) through by the multiplicative inverse of
$d_{j,j-k+m+1}$
modulo
$p^{(k-m)b}$
, we see that there is no loss in supposing that

Consequently, by taking suitable linear combinations of the congruences comprising (3.9), we discern that any solution of this system satisfies

in which

Note here that

Denote by
${\mathcal{D}}_{2}(\mathbf{u})$
the set of
${\mathcal{R}}(h-a)$
-equivalence classes of solutions of the system of congruences

with
$1\leqslant \mathbf{y}\leqslant p^{(k-m)b-a}$
satisfying (3.7). Then we have shown thus far that

Let
${\mathcal{D}}_{3}(\mathbf{v})$
denote the set of solutions of the congruence

with
$1\leqslant \mathbf{y}\leqslant p^{h-a}$
satisfying (3.7). For
$k-m\leqslant j\leqslant k$
, we have

Then we arrive at the upper bound

By combining (3.6), (3.12) and (3.13), we discern at this point that

Define the determinant

In view of (3.11), one has
$\unicode[STIX]{x1D713}_{k-m+l-1}^{\prime }(y_{i})\equiv ly_{i}^{l-1}~(\text{mod}~p)$
. It follows from (3.7) that

so that, since
$p>k$
, we have
$(J(\unicode[STIX]{x1D74D};\mathbf{y}),p)=1$
. We therefore deduce from [Reference Wooley13, Theorem 1], just as in the corresponding argument of the proof of [Reference Wooley17, Lemma 3.3] following [Reference Wooley17, equation (3.17)], that

and thus the conclusion of the lemma when
$a\geqslant 1$
follows at once from (3.2) and (3.14).
When
$a=0$
, we must apply some minor modifications to the above argument. In this case, we denote by
${\mathcal{D}}_{1}(\mathbf{n};\unicode[STIX]{x1D702})$
the set of
${\mathcal{R}}(h)$
-equivalence classes of solutions of the system of congruences (3.5) with
$1\leqslant \mathbf{z}\leqslant p^{kb}$
and
$\mathbf{z}\equiv \unicode[STIX]{x1D743}~(\text{mod}~p)$
for some
$\unicode[STIX]{x1D743}\in \unicode[STIX]{x1D6EF}_{0}^{m+1}(0)$
, and for which in addition
$z_{i}\not \equiv \unicode[STIX]{x1D702}~(\text{mod}~p)$
for
$1\leqslant i\leqslant m+1$
. Then, as in the opening paragraph of our proof, it follows from (3.1) that

But
$\text{card}({\mathcal{D}}_{1}(\mathbf{n};\unicode[STIX]{x1D702}))=\text{card}({\mathcal{D}}_{1}(\mathbf{n};0))$
, and
$\text{card}({\mathcal{D}}_{1}(\mathbf{n};0))$
counts the solutions of the system of congruences

with
$1\leqslant \mathbf{y}\leqslant p^{(k-m)b}$
satisfying (3.7), and in addition
$p\nmid y_{i}$
$(1\leqslant i\leqslant m+1)$
. Write

Then, since
$p>k$
, we have

We therefore conclude from [Reference Wooley13, Theorem 1] that

In view of (3.3), the conclusion of the lemma therefore follows from (3.15) when
$a=0$
. This completes our account of the proof of the lemma.◻
4 The conditioning process
The mean value
$K_{a,b}^{m,r}(X;\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
differs only slightly from the special case
$K_{a,b}^{k-1,k-1}(X;\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
considered in [Reference Wooley19], and thus our treatment of the conditioning process may be swiftly executed.
Lemma 4.1. Let
$a$
and
$b$
be integers with
$b>a\geqslant 1$
. Then, whenever
$s\geqslant 2r-1$
, one has

Proof. Our argument follows the proof of [Reference Wooley19, Lemma 4.1] with minor adjustments. Consider fixed integers
$\unicode[STIX]{x1D709}$
and
$\unicode[STIX]{x1D702}$
with
$\unicode[STIX]{x1D702}\not \equiv \unicode[STIX]{x1D709}~(\text{mod}~p)$
. Let
$T_{1}$
denote the number of integral solutions
$\mathbf{x}$
,
$\mathbf{y}$
,
$\mathbf{v}$
,
$\mathbf{w}$
of the system (2.13) counted by
$I_{a,b}^{r,r}(X;\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
in which
$v_{1},\ldots ,v_{s}$
together occupy at least
$r$
distinct residue classes modulo
$p^{b+1}$
, and let
$T_{2}$
denote the corresponding number of solutions in which these integers together occupy at most
$r-1$
distinct residue classes modulo
$p^{b+1}$
. Then

From (2.10) to (2.12), orthogonality and Schwarz’s inequality, one has

In order to treat
$T_{2}$
, we observe first that the hypothesis
$s\geqslant 2r-1$
ensures that there is an integer
$\unicode[STIX]{x1D701}\equiv \unicode[STIX]{x1D702}~(\text{mod}~p^{b})$
having the property that three at least of the variables
$v_{1},\ldots ,v_{s}$
are congruent to
$\unicode[STIX]{x1D701}$
modulo
$p^{b+1}$
. Hence, recalling the definitions (2.10) and (2.11), one finds from orthogonality and Hölder’s inequality that

By substituting (4.2) and (4.3) into (4.1), and recalling (2.16) and (2.17), we therefore conclude that

whence

This completes the proof of the lemma. ◻
Repeated application of Lemma 4.1 combines with a trivial bound for the mean value
$K_{a,b+H}^{r,r}(X)$
to deliver the basic conditioning lemma of this section.
Lemma 4.2. Let
$a$
and
$b$
be integers with
$1\leqslant a<b$
, and let
$H$
be any integer with
$H\geqslant 15$
. Suppose that
$b+H\leqslant (2\unicode[STIX]{x1D703})^{-1}$
. Then there exists an integer
$h$
with
$0\leqslant h<H$
having the property that

Proof. Repeated application of Lemma 4.1 yields the upper bound

On considering the underlying Diophantine systems, it follows from Lemma 2.3 that for each
$\unicode[STIX]{x1D709}$
and
$\unicode[STIX]{x1D702}$
, one has

Since
$M^{b+H}=(X^{\unicode[STIX]{x1D703}})^{b+H}\leqslant X^{1/2}$
, we deduce from (2.16) and (2.24) that

where

Since
$\unicode[STIX]{x1D706}\geqslant s+r$
, the lower bound
$b>a$
leads to the estimate

This completes the proof of the lemma, since with
$H\geqslant 15$
one has

◻
5 The precongruencing step
The alteration of the definition of
$K_{a,b}^{m,r}(X)$
relative to our earlier treatments does not prevent previous precongruencing arguments from applying, mutatis mutandis. However, we seek to ensure that our initial value of
$b$
is very large, so that subsequent iterations work efficiently. For this reason, the argument of the proof of [Reference Ford and Wooley4, Lemma 6.1] must be modified in several ways that demand a fairly complete account of the proof. We recall the definition (2.8) of
$B$
.
Lemma 5.1. There exist a prime number
$p$
, with
$M<p\leqslant 2M$
, and an integer
$h$
, with
$0\leqslant h\leqslant 4B$
, for which one has

Proof. The mean value
$J_{s+r}(X)$
counts the number of integral solutions of the system

with
$1\leqslant \mathbf{x},\mathbf{y}\leqslant X$
. Let
${\mathcal{P}}$
denote a set of
$\lceil (s+r)^{2}\unicode[STIX]{x1D703}^{-1}\rceil$
prime numbers lying in
$(M,2M]$
. That such a set exists is a consequence of the prime number theorem. The argument of the proof of [Reference Ford and Wooley4, Lemma 6.1] leading to equation (6.2) of that paper shows that for some
$p\in {\mathcal{P}}$
, one has
$J_{s+r}(X)\ll T(p)$
, where
$T(p)$
denotes the number of solutions of the system (5.1) counted by
$J_{s+r}(X)$
in which
$x_{1},\ldots ,x_{s+r}$
are distinct modulo
$p$
, and likewise
$y_{1},\ldots ,y_{s+r}$
are distinct modulo
$p$
.
We now examine the residue classes modulo
$p^{B}$
of a given solution
$\mathbf{x},\mathbf{y}$
counted by
$T(p)$
. Let
$\unicode[STIX]{x1D6C8}$
and
$\boldsymbol{\unicode[STIX]{x1D701}}$
be
$s$
-tuples with
$1\leqslant \unicode[STIX]{x1D6C8},\boldsymbol{\unicode[STIX]{x1D701}}\leqslant p^{B}$
satisfying the condition that for
$1\leqslant i\leqslant s$
, one has
$x_{i}\equiv \unicode[STIX]{x1D702}_{i}~(\text{mod}~p^{B})$
and
$y_{i}\equiv \unicode[STIX]{x1D701}_{i}~(\text{mod}~p^{B})$
. Recall the notation introduced prior to (2.18). Then, since
$x_{1},\ldots ,x_{s+r}$
are distinct modulo
$p$
, it follows that
$(x_{s+1},\ldots ,x_{s+r})\in \unicode[STIX]{x1D6EF}^{r}(\unicode[STIX]{x1D6C8})$
, and likewise one finds that
$(y_{s+1},\ldots ,y_{s+r})\in \unicode[STIX]{x1D6EF}^{r}(\boldsymbol{\unicode[STIX]{x1D701}})$
. Then, on considering the underlying Diophantine systems, one finds that

Write

Then, by applying Hölder’s inequality, and again considering the underlying Diophantine systems, we discern that

Hence, on recalling the definition (2.18), we obtain the upper bound

By modifying the argument of the proof of [Reference Ford and Wooley4, Lemma 6.1] leading from equation (6.3) to equation (6.6) of that paper, along the lines easily surmised from our proof of Lemma 4.1 above, one finds that

Iterating (5.3) in order to bound
$\widetilde{I}_{B}^{r,r}(X;\unicode[STIX]{x1D702})$
, just as in the argument concluding the proof of [Reference Ford and Wooley4, Lemma 6.1] (and see also (4.4) above), we discern from (5.2) that

By considering the underlying Diophantine systems, we deduce from Lemma 2.3 that

In this way, we find from (5.4) either that

for some index
$h$
with
$0\leqslant h\leqslant 4B$
, so that the conclusion of the lemma holds, or else that

In the latter case, since
$\unicode[STIX]{x1D706}\geqslant s+r$
, we obtain the upper bound

Invoking the definition (2.8) of
$\unicode[STIX]{x1D6FF}$
, we find that
$J_{s+r}(X)\ll X^{\unicode[STIX]{x1D706}-2\unicode[STIX]{x1D6FF}}$
, contradicting the lower bound (2.24) if
$X=X_{l}$
is large enough. We are therefore forced to accept the former upper bound (5.5), and hence the proof of the lemma is completed by reference to (2.20).◻
At this point, we fix the prime number
$p$
, once and for all, in accordance with Lemma 5.1.
6 The efficient congruencing step
Our strategy for executing the efficient congruencing process is based on that in our recent work [Reference Wooley19, §6], though in present circumstances only one phase is required, relating
$K_{a,b}^{m+1,r}(X)$
to
$K_{a,b}^{m,r}(X)$
and
$I_{b,b^{\prime }}^{r,r}(X)$
, for a suitable value of
$b^{\prime }$
. We begin our discussion of the efficient congruencing process with some additional notation. We define the generating function

in which we adopt the natural convention that
$\mathfrak{H}_{c,d}^{0}(\unicode[STIX]{x1D736};\unicode[STIX]{x1D709})=1$
. It is useful for future reference to record the consequence of Hölder’s inequality given in [Reference Wooley19, equation (6.2)], namely that whenever
$\unicode[STIX]{x1D714}$
is a real number with
$m\unicode[STIX]{x1D714}\geqslant 1$
, then

Lemma 6.1. Let
$m$
be an integer with
$0\leqslant m\leqslant r-1$
. Suppose that
$a$
and
$b$
are integers with
$0\leqslant a<b\leqslant \unicode[STIX]{x1D703}^{-1}$
and
$(k-m)b\geqslant (m+1)a$
, and suppose further that
$a\leqslant b/\sqrt{k}$
. Then, whenever
$b^{\prime }$
is an integer with
$a\leqslant b^{\prime }\leqslant (k-m)b-ma$
, one has

Proof. We first consider the situation in which
$a\geqslant 1$
. The argument associated with the case
$a=0$
is very similar, and so we are able to appeal later to a highly abbreviated argument for this case in order to complete the proof of the lemma. Consider fixed integers
$\unicode[STIX]{x1D709}$
and
$\unicode[STIX]{x1D702}$
with

The quantity
$K_{a,b}^{m+1,r}(X;\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
counts the number of integral solutions of the system (2.14) with
$m$
replaced by
$m+1$
, subject to the attendant conditions on
$\mathbf{x}$
,
$\mathbf{y}$
,
$\mathbf{u}$
,
$\mathbf{v}$
,
$\mathbf{w}$
,
$\mathbf{z}$
. Given such a solution of the system (2.14), the discussion leading to (2.15) shows that

In the notation introduced in §3, it follows that for some
$k$
-tuple of integers
$\mathbf{m}$
, both
$[\mathbf{x}~(\text{mod}~p^{kb})]$
and
$[\mathbf{y}~(\text{mod}~p^{kb})]$
lie in
${\mathcal{B}}_{a,b}^{m+1}(\mathbf{m};\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
. Write

Then, on considering the underlying Diophantine systems, we see from (2.12) and (6.4) that

where we write

We now partition the vectors in each set
${\mathcal{B}}_{a,b}^{m+1}(\mathbf{m};\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
into equivalence classes modulo
$p^{b^{\prime }}$
as in §3. Write
${\mathcal{C}}(\mathbf{m})={\mathcal{C}}_{a,b}^{m+1,b^{\prime }/b}(\mathbf{m};\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
. An application of Cauchy’s inequality leads via (3.2) and Lemma 3.1 to the bound

It may be helpful to note here that since
$b^{\prime }/b\leqslant k-m-ma/b$
, then in view of the relation (3.2), it follows from Lemma 3.1 that

Hence,

For each
$k$
-tuple
$\mathbf{m}$
and equivalence class
$\mathfrak{C}$
, the integral above counts solutions of (2.14) with the additional constraint that both
$[\mathbf{x}~(\text{mod}~p^{kb})]$
and
$[\mathbf{y}~(\text{mod}~p^{kb})]$
lie in
$\mathfrak{C}$
. In particular, one has
$\mathbf{x}\equiv \mathbf{y}~(\text{mod}~p^{b^{\prime }})$
. Moreover, the sets
${\mathcal{B}}_{a,b}^{m+1}(\mathbf{m};\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
are disjoint for distinct
$k$
-tuples
$\mathbf{m}$
with
$1\leqslant m_{j}\leqslant p^{jb}$
$(1\leqslant j\leqslant k)$
, so to each pair
$(\mathbf{x},\mathbf{y})$
there corresponds at most one pair
$(\mathbf{m},\mathfrak{C})$
. Thus, we deduce that

where
$H^{\dagger }(X;\unicode[STIX]{x1D709},\unicode[STIX]{x1D702})$
denotes the number of solutions of (2.14) subject to the additional condition
$\mathbf{x}\equiv \mathbf{y}~(\text{mod}~p^{b^{\prime }})$
. Hence, on considering the underlying Diophantine systems and recalling (6.1), we discern that

An inspection of the definition of
$\unicode[STIX]{x1D6EF}_{a}^{m}(\unicode[STIX]{x1D709})$
in the preamble to (2.10) reveals on this occasion that when
$\unicode[STIX]{x1D743}\in \unicode[STIX]{x1D6EF}_{a}^{m+1}(\unicode[STIX]{x1D709})$
, then

Then a further consideration of the underlying Diophantine systems leads from (6.6) via (6.1) to the upper bound

By applying Hölder’s inequality to the integral on the right-hand side of this relation, bearing in mind the definition (6.5), we obtain the bound

where

and



We now relate the mean values
$U_{i}$
to those introduced in §2. Observe first that a consideration of the underlying Diophantine systems leads from (6.9) via (6.1) and (2.10) to the upper bound

On recalling (2.12) and (2.17), we thus deduce that

Next, by employing (6.2) within (6.10) and (6.11), we find that

Notice here that since the condition (6.3) implies that
$\unicode[STIX]{x1D702}\not \equiv \unicode[STIX]{x1D709}~(\text{mod}~p)$
, and we have
$\unicode[STIX]{x1D701}\equiv \unicode[STIX]{x1D709}~(\text{mod}~p^{a})$
with
$a\geqslant 1$
, then once more one has
$\unicode[STIX]{x1D701}\not \equiv \unicode[STIX]{x1D702}~(\text{mod}~p)$
. In this way we deduce from (2.11) and (2.16) that

By substituting (6.12) and (6.13) into the relation

that is immediate from (6.7), and then recalling (6.8) and (2.17), the conclusion of the lemma follows when
$a\geqslant 1$
. When
$a=0$
, we must modify this argument slightly. In this case, from (2.19) and (2.20), we find that

The desired conclusion follows in this instance by pursuing the proof given above in the case
$a\geqslant 1$
, noting that the definition of
$\mathfrak{F}^{m+1}(\unicode[STIX]{x1D736};\unicode[STIX]{x1D702})$
ensures that the variables resulting from the congruencing argument avoid the congruence class
$\unicode[STIX]{x1D702}$
modulo
$p$
. This completes our account of the proof of the lemma.◻
We note that in our application of Lemma 6.1, we restrict to situations with
$0\leqslant m\leqslant r-1$
. Thus, since
$r\leqslant r_{0}$
, it follows from (2.2) that the hypothesis
$a\leqslant b/\sqrt{k}$
ensures that

Since
$b$
is large, we are therefore at liberty to apply Lemma 6.1 with a choice for
$b^{\prime }$
satisfying the condition
$b/b^{\prime }<1/\sqrt{k}$
, thereby preparing appropriately for subsequent applications of Lemma 6.1.
7 The multigrade combination
We next combine the estimates supplied by Lemma 6.1 so as to bound
$K_{a,b}^{r,r}(X)$
in terms of the mean values
$I_{b,k_{m}b}^{r,r}(X)$
$(0\leqslant m\leqslant r-1)$
, in which the integers
$k_{m}$
satisfy
$k_{m}\leqslant k-m-\lceil m/\sqrt{k}\rceil$
. Before announcing our basic asymptotic estimate, we define the exponents

In addition, we write

so that

Lemma 7.1. Suppose that
$a$
and
$b$
are integers with
$0\leqslant a<b\leqslant \unicode[STIX]{x1D703}^{-1}$
and
$(k-r+1)b\geqslant ra$
, and suppose further that
$a\leqslant b/\sqrt{k}$
. Then, whenever
$a^{\prime }$
is an integer with
$a^{\prime }\geqslant a$
for which
$(k-r+1)b\geqslant ra^{\prime }$
, one has

where we write
$b_{m}=(k-m)b-ma^{\prime }$
.
Proof. We prove by backwards induction that for
$0\leqslant l\leqslant r-1$
, one has

where

The conclusion of the lemma follows from the case
$l=0$
of (7.4), on noting that Lemma 2.2 delivers the estimate
$K_{a,b}^{0,r}(X)\ll J_{s+r}(X/M^{b})$
.
We observe first that the inductive hypothesis (7.4) holds when
$l=r-1$
, as a consequence of the case
$m=r-1$
of Lemma 6.1 and the definitions (7.1) and (7.2). Suppose then that
$J$
is a non-negative integer not exceeding
$r-2$
, and that the inductive hypothesis (7.4) holds for
$J<l\leqslant r-1$
. An application of Lemma 6.1 yields the estimate

On substituting this bound into the estimate (7.4) with
$l=J+1$
, one obtains the new upper bound

and thus the inductive hypothesis holds with
$l=J$
. This completes the inductive step, so in view of our earlier remarks, the conclusion of the lemma now follows.◻
We next convert Lemma 7.1 into a more portable form by making use of the anticipated magnitude operator
$\unicode[STIX]{x27E6}\,\cdot \,\unicode[STIX]{x27E7}$
introduced in (2.21) to (2.23).
Lemma 7.2. Suppose that
$a$
and
$b$
are integers with
$0\leqslant a<b\leqslant \unicode[STIX]{x1D703}^{-1}$
and
$(k-r+1)b\geqslant ra$
, and suppose further that
$a\leqslant b/\sqrt{k}$
. Then, whenever
$a^{\prime }$
is an integer with
$a^{\prime }\geqslant a$
for which
$(k-r+1)b\geqslant ra^{\prime }$
, one has

where
$b_{m}=(k-m)b-ma^{\prime }$
.
Proof. We find from Lemma 7.1 in combination with (2.23) that

where

On recalling (7.2) and (7.3), a modicum of computation confirms that

The conclusion of the lemma is therefore immediate from (7.5). ◻
We next turn to the task of establishing a multistep multigrade combination. Here, in order to keep complications under control, we discard some information not essential to our ultimate iterative process. We begin by introducing some additional notation. We recall that
$R$
is a positive integer sufficiently large in terms of
$s$
and
$k$
. We consider
$R$
-tuples of integers
$(m_{1},\ldots ,m_{R})\in [0,r-1]^{R}$
, to each of which we associate an
$R$
-tuple of integers
$\mathbf{h}(\mathbf{m})=(h_{1}(\mathbf{m}),\ldots ,h_{R}(\mathbf{m}))\in [0,\infty )^{R}$
. The integral tuples
$\mathbf{h}(\mathbf{m})$
will be fixed as the iteration proceeds, with
$h_{n}(\mathbf{m})$
depending at most on the first
$n$
coordinates of
$(m_{1},\ldots ,m_{R})$
. We may abuse notation in some circumstances by writing
$h_{n}(\mathbf{m},m_{n})$
or
$h_{n}(m_{1},\ldots ,m_{n-1},m_{n})$
in place of
$h_{n}(m_{1},\ldots ,m_{R})$
, reflecting the latter implicit dependence. We suppose that a (large) positive integer
$b$
has already been fixed. We then define the sequences
$(a_{n})=(a_{n}(\mathbf{m};\mathbf{h}))$
and
$(b_{n})=(b_{n}(\mathbf{m};\mathbf{h}))$
by putting

and then applying the iterative relations

Finally, we define the quantity
$\unicode[STIX]{x1D6E9}_{n}(\mathbf{m};\mathbf{h})$
for
$0\leqslant n\leqslant R$
by putting

Lemma 7.3. Suppose that
$a$
and
$b$
are integers with
$0\leqslant a<b\leqslant (16Rk^{2R}\unicode[STIX]{x1D703})^{-1}$
and
$(k-r+1)b\geqslant ra$
, and suppose further that
$a\leqslant b/\sqrt{k}$
. Then there exists a choice for
$\mathbf{h}(\mathbf{m})\in [0,\infty )^{R}$
, satisfying the condition that
$0\leqslant h_{n}(\mathbf{m})\leqslant 15k^{R}b$
$(1\leqslant n\leqslant R)$
, and for which one has

Proof. We prove by induction on
$l$
that when
$1\leqslant l\leqslant R$
, one has the upper bound

We observe first that, as a consequence of (7.3) and Lemma 7.2, one has

where

Moreover, it follows from Lemma 4.2 that for each
$m$
with
$0\leqslant m\leqslant r-1$
, there exists an integer
$h=h(m)$
, with
$0\leqslant h\leqslant 15k^{R}b$
, having the property that

whence from (2.22) and (2.23) we discern that

where

The inductive hypothesis (7.9) therefore follows in the case
$R=1$
by substituting (7.11) into (7.10). Notice here that we have discarded the factor
$M^{-hs/3}$
from (7.11), since at this stage of our argument it serves only as an unwelcome complication.
Suppose next that the inductive hypothesis (7.9) holds for
$1\leqslant l<L$
, with some integer
$L$
satisfying
$L\leqslant R$
. Consider the quantity
$\unicode[STIX]{x1D6E9}_{L-1}(\mathbf{m};\mathbf{h})$
for a given tuple
$\mathbf{m}\in [0,r-1]^{L-1}$
and a fixed tuple
$\mathbf{h}=\mathbf{h}(\mathbf{m})$
. We distinguish two possibilities. If it is the case that

then from (7.3) one finds that

so that

When

meanwhile, we apply Lemma 7.2 to obtain the bound

where

Two observations here serve to confirm that Lemma 7.2 is applicable. First, the hypothesis
$b\leqslant (16Rk^{2R}\unicode[STIX]{x1D703})^{-1}$
ensures that
$b_{L-1}\leqslant \unicode[STIX]{x1D703}^{-1}$
. Second, just as in the discussion concluding §6, an inductive argument shows that
$a_{n}\leqslant b_{n}/\sqrt{k}$
for each
$n$
, for the hypothesis
$1\leqslant r\leqslant r_{0}=k-\lceil 2\sqrt{k}\rceil +2$
ensures that whenever
$a_{n-1}\leqslant b_{n-1}/\sqrt{k}$
, then

whence
$b_{n}>\sqrt{k}b_{n-1}=\sqrt{k}a_{n}$
. The claimed relation therefore follows by induction from (7.6), and in particular one obtains

The applicability of Lemma 7.2 now confirmed, we again invoke Lemma 4.2, finding that for each integer
$m_{L}$
with
$0\leqslant m_{L}\leqslant r-1$
, there exists an integer
$h_{L}=h_{L}(\mathbf{m},m_{L})$
, with
$0\leqslant h_{L}\leqslant 15k^{R}b$
, having the property that

Thus, we deduce that

On substituting this estimate into (7.13), we deduce from (7.3) that

Applying this estimate in combination with (7.12) within the case
$l=L-1$
of the inductive hypothesis (7.9), we conclude that for some choice of the integer
$h_{L}=h_{L}(\mathbf{m})$
, one has the upper bound

This confirms the inductive hypothesis (7.9) for
$l=L$
, and thus the conclusion of the lemma follows by induction.◻
We remark that in obtaining the estimate (7.14), we made use of the trivial lower bound
$b_{L-1}\geqslant b$
. By discarding the power
$M^{b-b_{L-1}}$
at this point, we are throwing away potentially useful information. However, it transpires that the weak information made available by the factor
$M^{b}$
in the conclusion of Lemma 7.3 already suffices for our purposes in the main iteration.
8 The latent monograde process
Our objective in this section is to convert the block estimate encoded in Lemma 7.3 into a single monograde estimate that can be incorporated into our iterative method. We begin by recalling an elementary lemma from our previous work [Reference Wooley19].
Lemma 8.1. Suppose that
$z_{0},\ldots ,z_{l}\in \mathbb{C}$
, and that
$\unicode[STIX]{x1D6FD}_{i}$
and
$\unicode[STIX]{x1D6FE}_{i}$
are positive real numbers for
$0\leqslant i\leqslant l$
. Put
$\unicode[STIX]{x1D6FA}=\unicode[STIX]{x1D6FD}_{0}\unicode[STIX]{x1D6FE}_{0}+\cdots +\unicode[STIX]{x1D6FD}_{l}\unicode[STIX]{x1D6FE}_{l}$
. Then one has

Proof. This is [Reference Wooley19, Lemma 8.1]. ◻
Before proceeding further, we introduce some additional notation. Define the positive number
$s_{0}$
by means of the relation

in which
$\unicode[STIX]{x1D703}_{\pm }$
are defined as in (2.6). We recall that, in view of (2.7), one has
$s<s_{0}$
. Next, casting an eye towards the iterative relations (7.6) and (7.7), we define new sequences
$(\tilde{a}_{n})=(\tilde{a}_{n}(\mathbf{m}))$
and
$(\tilde{b}_{n})=(\tilde{b}_{n}(\mathbf{m}))$
by means of the relations

and

We then define

The motivation for defining the sequences
$(\tilde{a}_{n})$
and
$(\tilde{b}_{n})$
is to provide a base pair of sequences corresponding to the simplified situation in which
$h_{n}(\mathbf{m})=0$
for all
$n$
and
$\mathbf{m}$
. It transpires that the corresponding sequences
$(a_{n})$
and
$(b_{n})$
, equipped with potentially positive values of
$h_{n}(\mathbf{m})$
, may be bounded below by the sequences
$(\tilde{a}_{n})$
and
$(\tilde{b}_{n})$
, although this turns out to be less simple to establish than might be supposed.
Lemma 8.2. Suppose that
$\unicode[STIX]{x1D6EC}\geqslant 0$
, let
$a$
and
$b$
be integers with

and suppose further that
$a\leqslant b/\sqrt{k}$
. Suppose in addition that there are real numbers
$\unicode[STIX]{x1D713}$
,
$c$
and
$\unicode[STIX]{x1D6FE}$
, with

such that

Then, for some
$\mathbf{m}\in [0,r-1]^{R}$
, there are a real number
$h$
with
$0\leqslant h\leqslant 16k^{2R}b$
, and positive integers
$a^{\prime }$
and
$b^{\prime }$
with
$a^{\prime }\leqslant b^{\prime }/\sqrt{k}$
, such that

where
$\unicode[STIX]{x1D713}^{\prime }$
,
$c^{\prime }$
,
$\unicode[STIX]{x1D6FE}^{\prime }$
and
$b^{\prime }$
are real numbers satisfying the conditions

Moreover, the real number
$k_{\mathbf{m}}$
satisfies
$2^{R}\leqslant k_{\mathbf{m}}\leqslant k^{R}$
.
Proof. We deduce from the postulated bound (8.5) and Lemma 7.3 that there exists a choice of the tuple
$\mathbf{h}=\mathbf{h}(\mathbf{m})$
, with
$0\leqslant h_{n}(\mathbf{m})\leqslant 15k^{R}b$
$(1\leqslant n\leqslant R)$
, such that

Consequently, one has

Note that, in view of (7.3), one has

so that

Then we deduce from the definition (7.8) of
$\unicode[STIX]{x1D6E9}_{n}(\mathbf{m};\mathbf{h})$
that

In preparation for our application of Lemma 8.1, we examine the exponents
$\unicode[STIX]{x1D719}_{m_{1}}\cdots \unicode[STIX]{x1D719}_{m_{R}}$
occurring in the lower bound (8.8). Our plan is to apply Lemma 8.1 with exponents given by

in the natural sense. In order to analyze the quantity
$\unicode[STIX]{x1D6FA}$
that will emerge from the application of this lemma, we define

and then put
$\unicode[STIX]{x1D6FA}=B_{R}$
. From the iterative formulae (8.2) and (8.3), we obtain for
$n\geqslant 1$
the relations

and

We observe that from (7.1), one has

The summands on the right-hand side here may be rewritten in the shape

Hence, we obtain

in which
$\unicode[STIX]{x1D6E5}$
is defined via (2.3). In addition, one has

Utilizing the formulae just obtained, we conclude from (8.9) that for
$n\geqslant 1$
, one has

while from (8.7) and (8.10), one sees that

Then, on recalling (2.5), we arrive at the iterative relation

In addition, also from (2.5) and (8.2), one has the initial data

and, from (8.2) and (8.7), one finds that

Thus, we see also that

On recalling (2.6), one finds that the recurrence formula (8.11) has a solution of the shape

where, in view of (8.12) and (8.13), one has

and

Since
$\mathfrak{a}=\unicode[STIX]{x1D703}_{+}+\unicode[STIX]{x1D703}_{-}$
and
$\mathfrak{b}=\unicode[STIX]{x1D703}_{+}\unicode[STIX]{x1D703}_{-}$
, we therefore deduce that

In particular, on recalling (8.1), we find that
$s^{R}B_{R}=s_{0}^{R}$
, so that

Also, therefore, it follows from (2.7) that
$s<s_{0}$
and hence also that
$B_{R}>1$
.
Returning now to the application of Lemma 8.1, we note first that
$\unicode[STIX]{x1D6FA}=B_{R}$
and hence (8.8) yields the relation

But, in view of (8.4) and (8.14), one has
$\tilde{b}_{R}(\mathbf{m})/B_{R}=\unicode[STIX]{x1D70C}_{\mathbf{m}}$
, and thus we find that for some tuple
$\mathbf{m}\in [0,r-1]^{R}$
, one has

whence

Two further issues remain to be resolved, the first being the removal of the term
$M^{-3sk^{R}b}$
on the left-hand side of (8.15). We observe that the relations (8.3) ensure that
$\tilde{b}_{R}(\mathbf{m})\leqslant k^{R}$
, and hence (8.4) reveals that
$\unicode[STIX]{x1D70C}_{\mathbf{m}}<\tilde{b}_{R}(\mathbf{m})\leqslant k^{R}$
. Our hypothesis on
$c$
therefore ensures that

so that

Since

it follows from (8.15) that

Our final task consists of extracting appropriate constraints on the parameters
$a_{R}$
and
$b_{R}$
. We note first that from (7.6), one has
$a_{0}=\lfloor b/\sqrt{k}\rfloor \leqslant b/\sqrt{k}$
, so that on writing

and

we have

and

Define the sequences
$(a_{n}^{\ast })=(a_{n}^{\ast }(\mathbf{m};\mathbf{h}))$
and
$(b_{n}^{\ast })=(b_{n}^{\ast }(\mathbf{m};\mathbf{h}))$
by means of the relations

where

Then a comparison with (7.6) and (7.7) reveals that
$a_{n}^{\ast }=a_{n}$
and
$b_{n}^{\ast }=b_{n}$
for
$n\geqslant 1$
. Moreover, it is apparent from (8.2) and (8.3) that
$\tilde{a}_{n}b=a_{n}^{\ast }(\mathbf{m};\mathbf{0})$
and
$\tilde{b}_{n}b=b_{n}^{\ast }(\mathbf{m};\mathbf{0})$
for
$n\geqslant 0$
. We claim that for
$n\geqslant 0$
, one has

The validity of these inequalities for
$n=0$
is immediate from the definition of
$b_{n}^{\ast }(\mathbf{m};\mathbf{h})$
. We use an inductive argument to establish (8.17) for
$n\geqslant 1$
, though this requires some discussion.
Observe first that, by linearity, the recurrence sequence
$b_{n}^{\ast }(\mathbf{m};\mathbf{h})$
is given by the formula

in which
$c_{n,l}(\mathbf{m};\mathbf{h})$
is determined by the relations

with

We recall our assumption that
$r\leqslant k-\lceil 2\sqrt{k}\rceil +2$
, which ensures that

In view of this upper bound, we are able to prove by induction that

for each integer
$n$
with
$n\geqslant l-1$
. In order to confirm this assertion, note first that such is the case when
$n=l-1$
. Granted that
$c_{n,l}(\mathbf{m};\mathbf{h})>\sqrt{k}c_{n-1,l}(\mathbf{m};\mathbf{h})$
, meanwhile, one deduces from (8.19) that when
$n\geqslant l-1$
, one has

establishing this inductive hypothesis. Thus, in particular, one arrives at the lower bound
$c_{n,l}(\mathbf{m};\mathbf{h})\geqslant 0$
for all
$n$
. On substituting this conclusion into (8.18), we deduce that
$b_{n}^{\ast }(\mathbf{m};\mathbf{h})\geqslant b_{n}^{\ast }(\mathbf{m};\mathbf{0})$
for every
$n$
, confirming the first of the inequalities in (8.17).
Next, making use of the lower bound
$c_{n,l}(\mathbf{m};\mathbf{h})\geqslant 0$
, just obtained, within (8.19), we find that

so that

We therefore deduce from (8.18) and the bound
$h_{n}^{\ast }(\mathbf{m})\leqslant 16k^{R}b$
that

This confirms the second of the inequalities in (8.17).
We now make use of the inequalities (8.17). Observe first that in view of (8.4), one has

In addition, one has

Thus, there exists an integer
$h_{\mathbf{m}}$
, with
$0\leqslant h_{\mathbf{m}}\leqslant 16k^{2R}b$
, for which one has
$b_{R}=k_{\mathbf{m}}b+h_{\mathbf{m}}$
.
The argument confirming (8.21) shows also that for each
$n\geqslant 0$
, one has

Then it follows from (8.18) and (8.20) that

whence

Then we deduce from (8.16) that there exist integers
$b^{\prime }=b_{R}$
,
$a^{\prime }=a_{R}$
and
$h=h_{\mathbf{m}}$
satisfying the conditions

and for which

The desired conclusion (8.6) now follows, with all the associated conditions.
It now remains only to confirm the bounds on
$k_{\mathbf{m}}$
asserted in the final line of the statement of the lemma. For this, we note that by applying (8.22) in an inductive argument, one obtains from (8.4) the lower bound

Likewise, though more directly, one has

Thus,
$2^{R}\leqslant k_{\mathbf{m}}\leqslant k^{R}$
, and the proof of the lemma is complete.◻
9 The iterative process
We begin with a crude estimate of use at the conclusion of our argument.
Lemma 9.1. Suppose that
$a$
and
$b$
are integers with
$0\leqslant a<b\leqslant (2\unicode[STIX]{x1D703})^{-1}$
. Then, provided that
$\unicode[STIX]{x1D6EC}\geqslant 0$
, one has

Proof. On considering the underlying Diophantine equations, we deduce from Lemma 2.3 that

whence

This completes the proof of the lemma. ◻
We now come to the main event.
The Proof of Theorem 2.1.
We prove that when
$s$
is the largest integer smaller than
$s_{1}$
, then one has
$\unicode[STIX]{x1D6EC}\leqslant 0$
, for in such circumstances the conclusion of the lemma follows at once from (2.24). By reference to (2.5) and (2.6), we find that
$s<s_{1}=\unicode[STIX]{x1D703}_{+}$
. Thus, since
$R$
has been chosen sufficiently large in terms of
$s$
,
$k$
and
$\unicode[STIX]{x1D703}_{+}$
, the hypothesis
$s<s_{1}$
ensures that
$s<s_{0}$
, where
$s_{0}$
is defined by means of (8.1). Assume then that
$\unicode[STIX]{x1D6EC}\geqslant 0$
, for otherwise there is nothing to prove. We begin by noting that as a consequence of Lemma 5.1, one finds from (2.21) and (2.23) that there exists an integer
$h_{-1}$
with
$0\leqslant h_{-1}\leqslant 4B$
such that

We therefore deduce from (2.24) that

Next, we define sequences
$(\unicode[STIX]{x1D705}_{n})$
,
$(h_{n})$
,
$(a_{n})$
,
$(b_{n})$
,
$(c_{n})$
,
$(\unicode[STIX]{x1D713}_{n})$
and
$(\unicode[STIX]{x1D6FE}_{n})$
, for
$0\leqslant n\leqslant N$
, in such a way that

and

We note here that the sequences
$(a_{n})$
and
$(b_{n})$
are not directly related to our earlier use of these letters. Given a fixed choice for the sequences
$(a_{n})$
,
$(\unicode[STIX]{x1D705}_{n})$
and
$(h_{n})$
, the remaining sequences are defined by means of the relations




We put

so that both (9.2) and (9.3) hold with
$n=0$
as a consequence of our initial choice of
$\unicode[STIX]{x1D705}_{-1}$
and
$b_{-1}$
, together with (9.1). We prove by induction that for each non-negative integer
$n$
with
$n<N$
, the sequences
$(a_{m})_{m=0}^{n}$
,
$(\unicode[STIX]{x1D705}_{m})_{m=0}^{n}$
and
$(h_{m})_{m=-1}^{n}$
may be chosen in such a way that


and so that (9.2) and (9.3) both hold with
$n$
replaced by
$n+1$
.
Let
$0\leqslant n<N$
, and suppose that (9.2) and (9.3) both hold for the index
$n$
. We have already shown such to be the case for
$n=0$
. From (9.2) and (9.4), we find that
$b_{n}\leqslant 5(17k^{2R})^{n}B$
, whence, by invoking (2.8), we find that for
$0\leqslant n\leqslant N$
, one has

It is apparent from (9.5) and (9.6) that
$c_{n}$
and
$\unicode[STIX]{x1D713}_{n}$
are non-negative for all
$n$
. Observe also that since
$s\leqslant s_{0}$
and
$\unicode[STIX]{x1D705}_{m}\leqslant k^{R}$
, then by iterating (9.5) we obtain the bound

and by reference to (2.8) we see that
$c_{n}\leqslant (2\unicode[STIX]{x1D6FF})^{-1}\unicode[STIX]{x1D703}$
for
$0\leqslant n<N$
.
In order to bound
$\unicode[STIX]{x1D6FE}_{n}$
, we recall that
$s\leqslant s_{0}$
and iterate the relation (9.7) to deduce that

In addition, we find from (9.4) that for
$m\geqslant 0$
one has
$b_{m+1}\geqslant \unicode[STIX]{x1D705}_{m}b_{m}$
, so that an inductive argument yields the lower bound

Hence, we deduce from (9.11) that

Assembling this conclusion together with those of the previous paragraph, we have shown that (9.8) holds for
$0\leqslant n\leqslant N$
.
At this point in the argument, we may suppose that (9.3), (9.8) and (9.9) hold for the index
$n$
. An application of Lemma 8.2 therefore reveals that there exist real numbers
$\unicode[STIX]{x1D705}_{n}$
and
$h_{n}$
satisfying the constraints implied by (9.2) with
$n$
replaced by
$n+1$
, for which the upper bound (9.3) holds for some
$a_{n}$
with
$0\leqslant a_{n}\leqslant b_{n}/\sqrt{k}$
, also with
$n$
replaced by
$n+1$
. Our hypothesis on
$r$
, moreover, ensures that
$(k-r+1)b_{n+1}\geqslant ra_{n+1}$
. Hence (9.9) holds also with
$n$
replaced by
$n+1$
. This completes the inductive step, so that in particular the upper bound (9.3) holds for
$0\leqslant n\leqslant N$
.
We now exploit the bound just established. Since we have the upper bound
$b_{N}\leqslant 5(17k^{2R})^{N}B\leqslant (2\unicode[STIX]{x1D703})^{-1}$
, it is a consequence of Lemma 9.1 that

By combining this with (9.3) and (9.11), we obtain the bound

Meanwhile, an application of (9.10) in combination with (2.8) shows that
$X^{(c_{N}+1)\unicode[STIX]{x1D6FF}}<M$
. We therefore deduce from (9.12) that

Notice here that
$\unicode[STIX]{x1D705}_{n}\geqslant 2^{R}$
and

Hence,

so that

A further application of the lower bound
$b_{n}\geqslant \unicode[STIX]{x1D705}_{0}\cdots \unicode[STIX]{x1D705}_{n-1}B$
leads from (9.6) and the bound
$s\leqslant s_{0}$
to the relation

An inductive argument therefore delivers the lower bound

Thus, we deduce from (9.13) that

Since we are at liberty to take
$N$
as large as we please in terms of
$s$
and
$k$
, we are forced to conclude that
$\unicode[STIX]{x1D6EC}\leqslant 0$
. In view of our opening discussion, this completes the proof of the theorem.◻
10 Strongly diagonal behaviour
We are now equipped to establish Theorem 1.1 and its corollary. Let
$k$
be an integer with
$k\geqslant 7$
, put
$r=k-\lceil 2\sqrt{k}\rceil +2$
and suppose that
$s$
is a natural number with
$s\geqslant {\textstyle \frac{1}{4}}(k+1)^{2}$
. We note that [Reference Ford and Wooley4, Theorem 1.1] demonstrates that whenever
$1\leqslant s\leqslant {\textstyle \frac{1}{4}}(k+1)^{2}$
, then one has
$J_{s,k}(X)\ll X^{s+\unicode[STIX]{x1D700}}$
, and so we are certainly at liberty to assume that
$s>{\textstyle \frac{1}{4}}(k+1)^{2}$
. Write

and define
$s_{1}$
as in the statement of Theorem 2.1. We aim to show that the real number

satisfies the bounds

If these bounds be confirmed, then it follows from Theorem 2.1 that whenever

then
$J_{s,k}(X)\ll X^{s+\unicode[STIX]{x1D700}}$
, and the conclusion of Theorem 1.1 follows.
We observe in our first step that, in view of the lower bound
$s>{\textstyle \frac{1}{4}}(k+1)^{2}$
and our hypothesis
$k\geqslant 7$
, one has

Next, one finds that
$t_{0}\leqslant s_{1}$
provided only that

an inequality that is satisfied provided that

This last inequality amounts to the constraint
$kr-r(r-1)-r\geqslant 2\unicode[STIX]{x1D6E5}$
. We therefore conclude that
$t_{0}\leqslant s_{1}$
whenever
$r(k-r)\geqslant {\textstyle \frac{4}{3}}(k-6)$
. But when
$k\geqslant 7$
, it is easily verified that

and hence we confirm that
$t_{0}\leqslant s_{1}$
, as desired.
Meanwhile, the bound
$\unicode[STIX]{x1D6E5}<{\textstyle \frac{2}{3}}(k-6)$
ensures that

Since
$k\geqslant 7$
, moreover, one has

Then we have
$t_{0}\geqslant \max \{2r-1,\lceil {\textstyle \frac{1}{2}}r(r-1)+\unicode[STIX]{x1D6E5}\rceil \}$
, confirming the final claimed bound. In view of our earlier discussion, the proof of Theorem 1.1 is complete.
We turn now to the proof of Corollary 1.2. We suppose again that
$k\geqslant 7$
, and we put

Again defining
$\unicode[STIX]{x1D6E5}$
as in (10.1), in which
$r=k-\lceil 2\sqrt{k}\rceil +2$
, we find that

Hence, we deduce that

We therefore deduce from Theorem 1.1 that since

then
$J_{s,k}(X)\ll X^{s+\unicode[STIX]{x1D700}}$
. This completes the proof of the corollary.
11 A refinement of the core iteration
Our goal in this section is to outline an extension of our basic method that permits, for large exponents
$k$
, the refinement of Corollary 1.2 to deliver Theorem 1.3. The details of this extension are complicated enough that we aim for a somewhat abbreviated account, exploiting the discussion of §§2–10 so as to provide an outline.
In §2, we fixed
$r$
to be a parameter satisfying
$1\leqslant r\leqslant k-\lceil 2\sqrt{k}\rceil +2$
in order to ensure that, as the iteration bounding
$K_{a,b}^{r,r}(X)$
in terms of related mean values
$K_{a^{\prime },b^{\prime }}^{r,r}(X)$
proceeds, the exponents
$a^{\prime }$
and
$b^{\prime }$
satisfy the condition
$a^{\prime }\leqslant b^{\prime }/\sqrt{k}$
. The point of the latter constraint, in fact, is to ensure that the iteration tree of mean values does not encounter a situation in which
$b^{\prime }<a^{\prime }$
, which would be fatal for the method. By relaxing this condition, but pruning the iteration tree, we permit the possibility that
$a^{\prime }$
may be much closer to
$b^{\prime }$
in size, with the exponent mean
$s_{0}$
increased so that it becomes closer to
${\textstyle \frac{1}{2}}k(k+1)$
.
Let
$l=\lceil k^{1/3}\rceil$
and put
$r=k-l$
. We take
$s$
to be a natural number satisfying

We take
$R$
and
$N$
to be natural numbers sufficiently large in terms of
$s$
and
$k$
in the same manner as in §2, and define
$B$
,
$\unicode[STIX]{x1D703}$
and
$\unicode[STIX]{x1D6FF}$
as in (2.8). Other aspects of our initial set-up follow those of §2, mutatis mutandis. Lemma 6.1 remains valid provided that
$(k-m)b\geqslant (m+1)a$
. If instead one has
$(k-m)b<(m+1)a$
, then one may make use of an alternative inequality based on the application of Hölder’s inequality.
Lemma 11.1. Let
$m$
be an integer with
$0\leqslant m\leqslant r-1$
. Suppose that
$a$
and
$b$
are integers with
$0\leqslant a<b\leqslant \unicode[STIX]{x1D703}^{-1}$
. Then one has

where

Proof. On considering the underlying Diophantine systems, it follows from (2.12) and (2.14) that for some integers
$\unicode[STIX]{x1D709}$
and
$\unicode[STIX]{x1D702}$
, one has

Then, by Hölder’s inequality, we obtain

where

A further consideration of the underlying Diophantine systems reveals that
$I_{1}\leqslant I_{2}$
and
$I_{5}\leqslant I_{4}$
, and thus it follows from Lemma 2.2 and (2.12) that

The conclusion of the lemma follows at once. ◻
Our substitute for Lemma 7.2 must offer a surrogate for the upper bound supplied by Lemma 6.1 in those circumstances wherein
$m$
and
$a$
are both large. In this context, we introduce the parameter
$u=u(a,b)$
defined by

Notice that when
$0\leqslant m\leqslant u-1$
, one then has

We recall also the definitions of
$\unicode[STIX]{x1D719}_{m}$
and
$\unicode[STIX]{x1D719}^{\ast }$
from (7.1) and (7.2), and also write

Lemma 11.2. Suppose that
$a$
and
$b$
are integers with
$0\leqslant a<b\leqslant \unicode[STIX]{x1D703}^{-1}$
, and define
$u(a,b)$
by means of (11.1). Then, whenever
$a^{\prime }$
is an integer with
$a^{\prime }\geqslant a$
for which
$(k-r+1)b\geqslant ra^{\prime }$
, and
$u$
is an integer with
$0\leqslant u\leqslant u(a,b)$
, one has

where we write
$b_{m}=(k-m)b-ma^{\prime }$
.
Proof. By following the argument of the proof of Lemma 7.1, though substituting Lemma 11.1 in place of Lemma 6.1 when
$u\leqslant m\leqslant r-1$
, we deduce that

It may be useful here to note that

Thus, on recalling the definitions (2.21) to (2.23), we obtain the conclusion of the lemma, just as in the argument delivering Lemma 7.2. ◻
We next combine iterated applications of Lemma 11.2 so as to engineer a block multigrade process analogous to that delivered by Lemma 7.3. Before announcing the lemma that summarizes this process, we pause to modify the notation of §7. We consider
$R$
-tuples
$\mathbf{h}$
just as in the preamble to Lemma 7.3. The sequences
$(a_{n})=(a_{n}(\mathbf{m};\mathbf{h}))$
and
$(b_{n})=(b_{n}(\mathbf{m};\mathbf{h}))$
are now defined by putting

and then applying the iterative relations (7.7). We then define
$u_{n}(\mathbf{m};\mathbf{h})$
by putting

We emphasize here that, since
$a_{n-1}$
and
$b_{n-1}$
depend at most on the first
$n-1$
coordinates of
$\mathbf{m}$
and
$\mathbf{h}$
, then the same holds for
$u_{n}(\mathbf{m};\mathbf{h})$
. We use the notation

as shorthand for the ordered product

Finally, we redefine the quantity
$\unicode[STIX]{x1D6E9}_{n}(\mathbf{m};\mathbf{h})$
for
$0\leqslant n\leqslant R$
by putting

Lemma 11.3. Suppose that
$a$
and
$b$
are integers with

and suppose further that
$a\leqslant b/l$
. Suppose also that
$u_{1},\ldots ,u_{R}$
are integers with
$0\leqslant u_{i}\leqslant u_{i}(\mathbf{m};\mathbf{h})$
$(1\leqslant i\leqslant R)$
. Then there exists a choice for
$\mathbf{h}(\mathbf{m})\in [0,\infty )^{R}$
, satisfying the condition that
$0\leqslant h_{n}(\mathbf{m})\leqslant 15k^{R}b$
$(1\leqslant n\leqslant R)$
, and for which one has

Proof. One may follow the inductive strategy adopted in the proof of Lemma 7.3. The key difference in the present setting is that Lemma 11.2 contains additional factors in the estimate made available for
$\unicode[STIX]{x27E6}K_{a,b}^{r,r}(X)\unicode[STIX]{x27E7}$
. However, it follows from Lemma 11.2 that whenever
$u\leqslant u(a,b)$
, then

where

But we have

and hence
$2s(s-u)(s+r)\unicode[STIX]{x1D6FA}$
is equal to

Since we may suppose that
$s>r$
,
$s>u$
and
$s^{2}>ur$
, we find that
$\unicode[STIX]{x1D6FA}<0$
and hence

This relation serves as a substitute for Lemma 7.2 in the proof of Lemma 7.3, the proof of which now applies without serious modification. This completes our account of the proof of Lemma 11.3. ◻
We must now grapple with the somewhat daunting task of developing an analogue of Lemma 8.2 in the current setting. We now define the sequences
$(\tilde{a}_{n})=(\tilde{a}_{n}(\mathbf{m}))$
and
$(\tilde{b}_{n})=(\tilde{b}_{n}(\mathbf{m}))$
by putting

and then applying the relations (8.3). We define the parameters
$\tilde{u} _{i}=\tilde{u} _{i}(\mathbf{m})$
and
$\tilde{v}_{i}=\tilde{v}_{i}(\mathbf{m})$
for
$1\leqslant i\leqslant R$
by putting

Next, we define the positive real number
$s_{0}$
by means of the relation

and then put

We note here that the values of
$k_{\mathbf{m}}$
and
$\unicode[STIX]{x1D70C}_{\mathbf{m}}$
are not relevant when
$\mathbf{m}$
does not satisfy the condition
$0\leqslant m_{i}\leqslant \tilde{v}_{i}-1$
$(1\leqslant i\leqslant R)$
, and so we are indifferent to the values stemming from (11.6) in such circumstances.
Before proceeding further, we pause to relate
$u_{i}$
and
$\tilde{u} _{i}$
.
Lemma 11.4. Suppose that
$a\leqslant b/l$
and
$0\leqslant h_{n}(\mathbf{m})\leqslant 16k^{R}b$
$(1\leqslant n\leqslant R)$
. Then one has
$0\leqslant \tilde{u} _{n}(\mathbf{m})\leqslant u_{n}(\mathbf{m};\mathbf{h})$
$(1\leqslant n\leqslant R)$
.
Proof. We prove the asserted inequalities by induction, beginning with the case
$n=1$
. From (11.4), we find on the one hand that

whilst, on the other, from (11.2), one has

Thus
$\tilde{u} _{1}\leqslant u_{1}$
, confirming the inductive hypothesis when
$n=1$
. Suppose next that
$n\geqslant 1$
and
$\tilde{u} _{n}\leqslant u_{n}$
. One has

if and only if
$b_{n+1}/a_{n+1}\geqslant \tilde{b}_{n+1}/\tilde{a}_{n+1}$
, which is to say that

This lower bound is satisfied if and only if
$m_{n+1}\tilde{a}_{n}/\tilde{b}_{n}\geqslant (m_{n+1}a_{n}-h_{n+1})/b_{n}$
. The latter is automatically satisfied when either
$m_{n+1}=0$
or
$h_{n+1}\geqslant m_{n+1}a_{n}$
, and otherwise it is equivalent to the upper bound

However, when
$m_{n+1}>0$
, in view of our hypothesis
$\tilde{u} _{n}\leqslant u_{n}$
, one has

Thus, we conclude from (11.2), (11.4) and (11.7) that
$\tilde{u} _{n+1}\leqslant u_{n+1}$
. This confirms the inductive hypothesis and hence we deduce that
$0\leqslant \tilde{u} _{n}\leqslant u_{n}$
for
$1\leqslant n\leqslant R$
, as claimed.◻
We next introduce some additional quantities of use in our ultimate application of Lemma 8.1. Define the exponents
$\unicode[STIX]{x1D6FD}_{\mathbf{m}}^{(l)}$
and
$\unicode[STIX]{x1D6FE}_{\mathbf{m}}^{(l)}$
just as in the discussion following (8.8) above. We now put

In order to understand these sequences, we introduce some auxiliary sequences
$\tilde{A}_{n}$
and
$\tilde{B}_{n}$
as follows. We put

The sequence
$\tilde{B}_{n}$
is then defined for
$n\geqslant 1$
by means of the relations

and

Finally, we put

Lemma 11.5. When
$n\geqslant 1$
, one has
$B_{n}\geqslant \tilde{B}_{n}$
. In particular, one has the lower bound
$B_{R}\geqslant \tilde{B}_{R}$
and hence

Proof. We have

The innermost sum here is

where

We observe that when
$u$
is a positive integer, one has

whence

Thus, we conclude that

One finds in like manner that

Write

Then we see from (11.15) that
$s\mathfrak{B}(v)$
is equal to

Also, from (11.16), we find that
$s\mathfrak{A}(v)$
is equal to

On substituting these formulae into (11.14), we find that

where

and

We next seek to estimate the expression
$\unicode[STIX]{x1D6F6}$
, this requiring us to obtain an upper bound for
$h$
. We begin with a proof that
$\tilde{v}_{n}(\mathbf{m})\geqslant k(1-2/l)$
for
$1\leqslant n\leqslant R$
. Note first that, since we assume
$k$
to be sufficiently large and
$l=\lceil k^{1/3}\rceil$
, one finds from (11.4) that

Meanwhile, when
$n\geqslant 1$
, the condition

that follows from (11.4) ensures that

whence

We therefore deduce that
$\tilde{v}_{n}\geqslant k(1-2/l)$
for
$n\geqslant 1$
, as we had claimed. It follows, in particular, that

We use this opportunity also to recall the bounds
$s\geqslant {\textstyle \frac{1}{2}}k(k+1)-4k$
and
$l\leqslant k^{1/3}+1$
.
Before exploiting the crude bound for
$h$
just obtained, we make use of the more opaque though stronger bound

On substituting these bounds into (11.18), we see that

Finally, we observe that

By combining these estimates with (11.17), we arrive at the relation

Then we deduce from (11.13) that

On the other hand, we have

whence

By combining (11.19) and (11.20), we conclude at this point that

In addition, we have the initial data

and hence

Our goal is now to extract from the recurrence inequality (11.21) a lower bound for
$B_{n}$
. Were we to be presented with an equation, this would be straightforward, but in present circumstances we must work less directly by relating
$B_{n}$
to
$\tilde{B}_{n}$
. By reference to (11.11), we see that
$\tilde{B}_{n}$
satisfies a recurrence equation related to the inequality (11.19).
Our first observation is that the recurrence formula (11.11) has a solution of the shape

where

and

Then we deduce that

Note that since
$k$
is large, it follows from (11.9) and (11.12) that

and

Thus, in particular, we deduce from (11.22) that
$\tilde{B}_{n}$
is positive for every natural number
$n$
.
We now reinterpret the recurrence inequality (11.21) as a recurrence equation with variable shifts. Let
$(g_{n})$
be a sequence of positive real numbers. Then, for an appropriate choice of this sequence
$(g_{n})$
, the recurrence sequence
$B_{n}$
may be interpreted as the solution of the new recurrence sequence

with the initial data

Define the real numbers
$\tilde{g}_{n}$
by means of the relation

Thus, in particular, it follows from (11.10) and (11.25) that

and from (11.11) and (11.24) that

We claim that
$\tilde{g}_{n}\geqslant 0$
, whence also
$B_{n}\geqslant \tilde{B}_{n}$
, for each natural number
$n$
.
In order to verify the last claim, we prove by induction that for each natural number
$n$
, one has

We first confirm this inductive hypothesis for
$n=1$
. Here we note that when
$\tilde{g}_{1}=g_{1}=0$
, then the trivial lower bound
$\tilde{g}_{2}=g_{2}\geqslant 0$
that follows from (11.27) suffices to confirm (11.29). When
$g_{1}\neq 0$
, meanwhile, we find from (11.10) and (11.27) that

The positivity of
$g_{1}$
,
$g_{2}$
and

consequently ensures that

This confirms the inductive hypothesis when
$n=1$
.
Suppose next that
$n\geqslant 2$
and

Then, on making use of (11.11) and (11.28), one deduces that

This confirms the inductive hypothesis with
$n+1$
in place of
$n$
, and so by applying induction we deduce that (11.29) holds for every natural number
$n$
. A particular consequence of the lower bound (11.29) is that
$\tilde{g}_{n}\geqslant 0$
for every natural number
$n$
. Finally, we conclude from (11.26) that
$B_{n}\geqslant \tilde{B}_{n}$
for every natural number
$n$
.
In order to complete the proof of the lemma, we note first from (11.5) and (11.8) that
$s_{0}^{R}=s^{R}B_{R}$
. Since
$B_{R}\geqslant \tilde{B}_{R}$
, the final conclusion of the lemma follows from (11.22).◻
We next turn to the problem of controlling the behaviour of
$a_{R}$
and
$b_{R}$
.
Lemma 11.6. Suppose that the tuples
$\mathbf{m}$
and
$\mathbf{h}(\mathbf{m})$
satisfy the conditions

Then one has

and furthermore

Proof. We again make use of auxiliary recurrence sequences in order to disentangle information on recurrence inequalities. Write

and

Then we find that

and

We define the sequences
$(a_{n}^{\ast })=(a_{n}^{\ast }(\mathbf{m};\mathbf{h}))$
and
$(b_{n}^{\ast })=(b_{n}^{\ast }(\mathbf{m};\mathbf{h}))$
by means of the relations

and

It follows that
$a_{n}=a_{n}^{\ast }$
and
$b_{n}=b_{n}^{\ast }$
for
$n\geqslant 1$
. We next put

and we seek to show in the first instance that
$g_{n}^{\ast }\geqslant 0$
for each
$n$
. This of course implies in particular that
$b_{R}^{\ast }\geqslant \tilde{b}_{R}b=k_{\mathbf{m}}b$
.
We prove that
$g_{n}^{\ast }\geqslant 0$
for each
$n$
by induction. Observe first that by applying (11.30) in combination with the recurrence relations for
$\tilde{a}_{n}$
,
$\tilde{b}_{n}$
,
$a_{n}^{\ast }$
,
$b_{n}^{\ast }$
, we obtain

whence
$g_{0}^{\ast }\tilde{b}_{1}-\tilde{b}_{0}g_{1}^{\ast }=-h_{1}^{\ast }\leqslant 0$
. Moreover, if we assume that

then we find in like manner that

Thus, we obtain the bound

We thus conclude from (11.31) that
$g_{n}^{\ast }\tilde{b}_{n+1}-\tilde{b}_{n}g_{n+1}^{\ast }\leqslant 0$
, thereby confirming the inductive hypothesis (11.31) with
$n+1$
in place of
$n$
. We therefore deduce by induction that
$g_{n-1}^{\ast }\tilde{b}_{n}\leqslant \tilde{b}_{n-1}g_{n}^{\ast }$
for every
$n$
, whence

In this way, we therefore conclude that
$b_{R}=b_{R}^{\ast }\geqslant \tilde{b}_{R}b=k_{\mathbf{m}}b$
, as desired.
Having confirmed the lower bound on
$b_{R}(\mathbf{m};\mathbf{h})$
claimed in the statement of the lemma, we turn our attention next to the upper bound. Observe that the recurrence relations for
$\tilde{a}_{n}$
,
$\tilde{b}_{n}$
,
$a_{n}^{\ast }$
,
$b_{n}^{\ast }$
lead us from (11.30) to the relations

and

In particular, we have the trivial upper bound

and so we deduce by induction that

Hence, on recalling (11.30) once again, we obtain the bound

Collecting together the conclusions of the last two paragraphs, we find that

thereby confirming the first assertion of the lemma. It remains now only to bound
$a_{R}$
in terms of
$b_{R}$
. To this end, we observe that for each
$n\geqslant 0$
, it follows from (11.4) and our hypothesis
$m_{n+1}\leqslant \tilde{u} _{n}(\mathbf{m})-1$
that
$m_{n+1}\leqslant r\tilde{b}_{n}/(\tilde{a}_{n}+\tilde{b}_{n})$
and hence

We therefore deduce from (11.32) that
$g_{n+1}^{\ast }\geqslant lg_{n}^{\ast }$
for
$n\geqslant 0$
. Consequently, recalling again the relation (11.30), we see that

We have therefore shown that
$a_{R}\leqslant b_{R}/l$
, as desired. This completes the proof of the lemma.◻
Before initiating our discussion of the next lemma, we observe that in view of (11.23) and our hypotheses concerning
$s$
, one has

Then it follows that, provided
$R$
is sufficiently large in terms of
$s$
and
$k$
, one has
$s<\unicode[STIX]{x1D703}_{+}$
, and so we deduce from Lemma 11.5 that

Lemma 11.7. Suppose that
$\unicode[STIX]{x1D6EC}\geqslant 0$
, let
$a$
and
$b$
be integers with
$0\leqslant a<b\leqslant (20Rk^{2R}\unicode[STIX]{x1D703})^{-1}$
and suppose further that
$a\leqslant b/l$
. Suppose in addition that there are real numbers
$\unicode[STIX]{x1D713}$
,
$c$
and
$\unicode[STIX]{x1D6FE}$
, with

such that

Then, for some
$\mathbf{m}\in [0,r-1]^{R}$
, there are a real number
$h$
with
$0\leqslant h\leqslant 16k^{2R}b$
and positive integers
$a^{\prime }$
and
$b^{\prime }$
with
$a^{\prime }\leqslant b^{\prime }/l$
, such that

where
$\unicode[STIX]{x1D713}^{\prime }$
,
$c^{\prime }$
,
$\unicode[STIX]{x1D6FE}^{\prime }$
and
$b^{\prime }$
are real numbers satisfying the conditions

Moreover, the real number
$k_{\mathbf{m}}$
satisfies
$2^{R}\leqslant k_{\mathbf{m}}\leqslant k^{R}$
.
Proof. We deduce from the postulated bound (11.34), together with Lemmas 11.3 and 11.4, that there exists a choice
$\mathbf{h}=\mathbf{h}(\mathbf{m})$
of tuples, with

such that

Consequently, one has

Note that from (7.1) one has

so that

Then we deduce from the definition (11.3) of
$\unicode[STIX]{x1D6E9}_{n}(\mathbf{m};\mathbf{h})$
that

Put

so that in view of Lemma 11.5 one has
$\unicode[STIX]{x1D6FA}\geqslant \tilde{B}_{R}$
. Then an application of Lemma 8.1 to (11.36) yields the relation

We see from (11.5) and (11.37) that
$\unicode[STIX]{x1D6FA}=(s_{0}/s)^{R}$
, and so it follows from (11.6) that
$\unicode[STIX]{x1D6FA}/\tilde{b}_{R}(\mathbf{m})=1/\unicode[STIX]{x1D70C}_{\mathbf{m}}$
. Thus, we conclude that for some
$R$
-tuple
$\mathbf{m}$
, one has

We must again handle the removal of the term
$M^{-3sk^{R}b}$
on the left-hand side of (11.38). We note that, just as in the concluding paragraph of the proof of Lemma 11.6, one has
$\tilde{b}_{n+1}\geqslant l\tilde{b}_{n}$
$(n\geqslant 0)$
and hence
$\tilde{b}_{n}(\mathbf{m})\geqslant 0$
. Hence, a direct induction leads from (8.3) to the upper bound
$\tilde{b}_{R}(\mathbf{m})\leqslant k^{R}$
, and we find by means of (11.6) and (11.33) that

We may therefore follow the analysis leading to (8.16), mutatis mutandis, to conclude as before that

We find from Lemma 11.6 that one has
$a_{R}\leqslant b_{R}/l$
, and that

Then we conclude that there exist integers
$b^{\prime }=b_{R}$
,
$a^{\prime }=a_{R}$
and
$h$
satisfying the conditions

for which

The desired conclusion (11.35) follows at once, together with its associated conditions.
The only task that remains is to confirm the bounds
$2^{R}\leqslant k_{\mathbf{m}}\leqslant k^{R}$
. On the one hand, just as in the proof of Lemma 11.6, one has
$\tilde{b}_{n+1}\geqslant l\tilde{b}_{n}$
for each
$n\geqslant 0$
. Then an inductive argument confirms that
$k_{\mathbf{m}}=\tilde{b}_{R}\geqslant l^{R}\tilde{b}_{0}\geqslant 2^{R}$
. On the other hand, we have already shown that
$\tilde{b}_{R}\leqslant k^{R}$
and thus
$k_{\mathbf{m}}=\tilde{b}_{R}\leqslant k^{R}$
. This completes the proof of the lemma.◻
We now employ the conclusion of the last lemma to bound
$J_{s+r}(X)$
.
Theorem 11.8. Suppose that
$s$
,
$k$
and
$r$
are natural numbers with
$k$
sufficiently large,
$r=k-\lceil k^{1/3}\rceil$
and
$1\leqslant s\leqslant s_{1}$
, where

Then, for each
$\unicode[STIX]{x1D700}>0$
, one has
$J_{s+r}(X)\ll X^{s+r+\unicode[STIX]{x1D700}}$
.
Proof. Defining
$s_{0}$
via (11.5) as in the preamble to Lemma 11.4, one finds that there are no serious modifications required in order to apply the argument of the proof of Theorem 2.1 to pass from Lemma 11.7 to the conclusion of the theorem, but with
$s_{0}$
in place of
$s_{1}=\unicode[STIX]{x1D703}_{+}$
. The argument leading to (11.33) shows, however, that whenever
$s<s_{1}$
, one has also that
$s<s_{0}$
. The conclusion of the theorem is now immediate.◻
In order to establish Theorem 1.3, we have only to note that, on recalling (11.23), one finds that

Thus, whenever

one has
$J_{s,k}(X)\ll X^{s+\unicode[STIX]{x1D700}}$
.
12 Some consequences of Theorem 1.3
We finish this paper by turning our attention to a few consequences of the asymptotically sharpest of our conclusions, namely Theorem 1.3. We begin by proving Theorem 1.5. Suppose that
$k$
is a sufficiently large natural number and put

In addition, take
$s={\textstyle \frac{1}{2}}k(k+1)$
and put
$\unicode[STIX]{x1D6E5}=s-\mathfrak{s}$
. Then we find that

Consequently, on making use of the trivial estimate
$f_{k}(\unicode[STIX]{x1D736};X)=O(X)$
, we deduce from (2.1) in combination with Theorem 1.3 that

The conclusion of Theorem 1.5 is now immediate.
We next point out an application of Theorem 1.3 to Tarry’s problem. Recall the definition of the natural number
$W(k,h)$
associated with the system of equations (1.5).
Theorem 12.1. When
$h$
and
$k$
are natural numbers with
$h\geqslant 2$
and
$k$
sufficiently large, one has
$W(k,h)\leqslant {\textstyle \frac{1}{2}}k(k+1)+1$
.
Proof. The argument of the proof of [Reference Wooley15, Theorem 1.3] shows that
$W(k,h)\leqslant s$
whenever one can establish the estimate

But, as a consequence of Theorem 1.3, when
$k$
is sufficiently large and

one has
$J_{s,k+1}(X)\ll X^{s+\unicode[STIX]{x1D700}}$
. Observe, however, that
$X^{s+\unicode[STIX]{x1D700}}=o(X^{2s-k(k+1)/2})$
provided only that
$s<2s-{\textstyle \frac{1}{2}}k(k+1)$
, which is to say that
$s\geqslant {\textstyle \frac{1}{2}}k(k+1)+1$
. Moreover, when
$k$
is sufficiently large, one finds that the value
$s={\textstyle \frac{1}{2}}k(k+1)+1$
satisfies the condition (12.2) and thus we have
$W(k,h)\leqslant {\textstyle \frac{1}{2}}k(k+1)+1$
. This completes the proof of the theorem.◻
The problem of estimating
$W(k,h)$
has been investigated extensively by Wright and Hua (see [Reference Hua6], [Reference Hua7] and [Reference Wright21]). Hua was able to show that
$W(k,h)\leqslant k^{2}(\log k+O(1))$
. This bound was improved in [Reference Wooley15, Theorem 1.3] by means of the efficient congruencing method, delivering the upper bound
$W(k,h)\leqslant k^{2}+k-2$
. In our most recent work on multigrade efficient congruencing, this bound was further improved in [Reference Wooley19, Theorem 12.1] to obtain
$W(k,h)\leqslant {\textstyle \frac{5}{8}}(k+1)^{2}$
for
$k\geqslant 3$
. The bound recorded in Theorem 12.1 above achieves the limit of the methods currently employed in which an elementary application of Vinogradov’s mean value theorem is applied, as described in [Reference Wooley14] and enhanced in [Reference Wooley19, §12]. As far as the author is aware, there is no lower bound available on
$W(k,h)$
superior to
$W(k,h)\geqslant k+1$
. There consequently remains a substantial gap between our upper and lower bounds in Tarry’s problem.
Finally, we explore the consequences of Theorem 1.3 in the context of Waring’s problem. When
$s$
and
$k$
are natural numbers, let
$R_{s,k}(n)$
denote the number of representations of the natural number
$n$
as the sum of
$s$
$k$
th powers of positive integers. A formal application of the circle method suggests that for
$k\geqslant 3$
and
$s\geqslant k+1$
, one should have

where

With suitable congruence conditions imposed on
$n$
, one has
$1\ll \mathfrak{S}_{s,k}(n)\ll n^{\unicode[STIX]{x1D700}}$
, so that the conjectured relation (12.3) constitutes an honest asymptotic formula. Let
$\widetilde{G}(k)$
denote the least integer
$t$
with the property that, for all
$s\geqslant t$
, and all sufficiently large natural numbers
$n$
, one has the asymptotic formula (12.3). By combining the conclusion of Theorem 1.3 with our recent work concerning the asymptotic formula in Waring’s problem [Reference Wooley16], and the enhancement [Reference Ford and Wooley4, Theorem 8.5] of Ford’s work [Reference Ford3], we are able to derive new upper bounds for
$\widetilde{G}(k)$
when
$k$
is sufficiently large.
Theorem 12.2. Let
$\unicode[STIX]{x1D709}$
denote the real root of the polynomial
$6\unicode[STIX]{x1D709}^{3}+3\unicode[STIX]{x1D709}^{2}-1$
and put
$C=(5+6\unicode[STIX]{x1D709}-3\unicode[STIX]{x1D709}^{2})/(2+6\unicode[STIX]{x1D709})$
, so that

Then, for large values of
$k$
, one has
$\widetilde{G}(k)<Ck^{2}+O(k^{5/3})$
.
For comparison, in [Reference Wooley19, Theorem 1.3], we derived an analogous bound in which
$C$
is replaced by the slightly larger number
$1.542\,749\ldots .$
The Proof of Theorem 12.2.
In general, we write
$\unicode[STIX]{x1D6E5}_{s,k}$
for the least real number with the property that, for all
$\unicode[STIX]{x1D700}>0$
, one has

Then it follows from [Reference Wooley19, Lemma 10.7] that one has
$\widetilde{G}(k)\leqslant \lfloor u_{1}(k)\rfloor +1$
, where

and

Suppose that
$k$
is a large natural number and let
$\unicode[STIX]{x1D6FD}$
be a positive parameter to be determined in due course. We take

Then it follows from Theorem 1.3, just as in the discussion leading to (12.1), that the exponent
$\unicode[STIX]{x1D6E5}_{v,k}$
is permissible, where

Also, by taking
$t=k^{2}-k+1$
, one sees from [Reference Wooley19, Corollary 1.2] that
$\unicode[STIX]{x1D6E5}_{t,k}=0$
. Then we deduce that

It follows that

A modest computation confirms that the optimal choice for the parameter
$\unicode[STIX]{x1D6FD}$
is
$\unicode[STIX]{x1D709}$
, where
$\unicode[STIX]{x1D709}$
is the real root of the polynomial equation
$6\unicode[STIX]{x1D709}^{3}+3\unicode[STIX]{x1D709}^{2}-1=0$
. With this choice for
$\unicode[STIX]{x1D6FD}$
, one finds that

The conclusion of Theorem 12.2 is now immediate. ◻