0808.0125/13.tex
1: \documentclass[11pt, a4paper, reqno, english]{smfart}
2: 
3: \usepackage[margin=3.5cm]{geometry} 
4: \usepackage{amsfonts, amsthm, amsmath, amscd, amssymb}
5: 
6: 
7: \renewcommand\AA{\mathbb{A}}
8: \newcommand\A{\mathcal{A}}
9: \newcommand\QQ{\mathbb{Q}}
10: \newcommand\cQ{\overline{\mathbb{Q}}}
11: \newcommand\CC{\mathbb{C}}
12: \newcommand\NN{\mathbb{N}}
13: \newcommand\RR{\mathbb{R}}
14: \newcommand\MM{\mathcal{M}}
15: \newcommand\WW{\mathcal{W}}
16: \newcommand\R{\mathcal{R}}
17: \newcommand\B{\mathcal{B}}
18: \newcommand\TT{\mathcal{T}}
19: \newcommand\EE{\mathcal{E}}
20: \newcommand\ZZ{\mathbb{Z}}
21: \newcommand\ZZp{\ZZ_{>0}}
22: \newcommand\ZZnz{\ZZ_{\ne 0}}
23: \newcommand\PP{\mathbb{P}}
24: \newcommand\x{\mathbf{x}}
25: \newcommand\y{\mathbf{y}}
26: \renewcommand\v{\mathbf{v}}
27: \newcommand\z{\mathbf{z}}
28: \renewcommand\k{\mathbf{k}}
29: \renewcommand\d{\,\mathrm{d}}
30: \newcommand{\congr}[3]{{#1} \equiv {#2}\ (\mathrm{mod}\ {#3})}
31: \DeclareMathOperator{\hcf}{gcd}
32: \DeclareMathOperator{\Pic}{Pic}
33: \DeclareMathOperator{\rank}{rank}
34: \DeclareMathOperator{\Res}{Res}
35: \DeclareMathOperator{\vol}{vol}
36: \DeclareMathOperator{\Spec}{Spec}
37: \DeclareMathOperator{\disc}{Disc}
38: \DeclareMathOperator{\supp}{supp}
39: \DeclareMathOperator{\sing}{sing}
40: \DeclareMathOperator{\meas}{meas}
41: \DeclareMathOperator{\Gal}{Gal}
42: \DeclareMathOperator{\Tr}{Tr}
43: \DeclareMathOperator{\Norm}{N}
44: 
45: 
46: 
47: \newcommand{\cp}[2]{{\hcf(#1,#2)=1}}
48: \newcommand{\ncp}[2]{{\hcf(#1,#2) > 1}}
49: \newcommand{\Aone}{{\mathbf A}_1}
50: \newcommand{\Atwo}{{\mathbf A}_2}
51: \newcommand{\Athree}{{\mathbf A}_3}
52: \newcommand{\Afour}{{\mathbf A}_4}
53: \newcommand{\Afive}{{\mathbf A}_5}
54: \newcommand{\Dfour}{{\mathbf D}_4}
55: \newcommand{\Dfive}{{\mathbf D}_5}
56: \newcommand{\Esix}{{\mathbf E}_6}
57: \newcommand{\tS}{{\widetilde S}}
58: \newcommand{\tM}{{\widetilde M}}
59: \newcommand{\tX}{{\widetilde X}}
60: \renewcommand{\le}{\leqslant}
61: \renewcommand{\ge}{\geqslant}
62: \renewcommand{\leq}{\leqslant}
63: \renewcommand{\geq}{\geqslant}
64: \newcommand{\bet}{\boldsymbol{\eta}}
65: \newcommand{\bla}{\boldsymbol{\lambda}}
66: \newcommand{\bal}{\boldsymbol{\alpha}}
67: \newcommand{\bxi}{\boldsymbol{\xi}}
68: \newcommand{\ex}[1]{*+<10pt>[o][F]{#1}}
69: \newcommand\rto{\dashrightarrow}
70: \newcommand\e{\eta}
71: \newcommand\ve{\varepsilon}
72: \newcommand\RE{\Re e}
73: \newcommand\phis{\phi^*}
74: \newcommand\phid{\phi^\dagger}
75: \newcommand\la{\lambda}
76: \newcommand\al{\alpha}
77: \newcommand\be{\beta}
78: \newcommand\D{\Delta}
79: 
80: \newcommand\sfl{\mathsf{\Lambda}}
81: 
82: \newtheorem{theorem}{Theorem}
83: \newtheorem*{theorem-f}{Th\'eor\`eme}
84: \newtheorem*{cor}{Corollary}
85: \newtheorem*{cor-f}{Corollaire}
86: \newtheorem*{con}{Conjecture}
87: \newtheorem{lemma}{Lemma}
88: \newtheorem{proposition}{Proposition}
89: \newtheorem{conj}{Conjecture}
90: \theoremstyle{definition}
91: \newtheorem{remark}{Remarque}
92: \newtheorem{example}{Example}
93: \newtheorem*{ack}{Acknowledgements}
94: \newtheorem*{notat}{Notation}
95: 
96: \newcommand{\odd}[2]{{(#1,#2)_\flat}}
97: \newcommand{\M}{\mathfrak{M}}
98: \renewcommand{\ss}{\mathfrak{S}}
99: \newcommand{\ii}{\mathfrak{I}}
100: \newcommand{\n}{\mathfrak{n}}
101: \newcommand{\m}{\mathfrak{m}}
102: \newcommand{\Xns}{X_{\mathrm{ns}}}
103: 
104: 
105: \numberwithin{equation}{section}
106: 
107: 
108: %%%CT
109: \newcommand\kbar{{\overline k}}
110: 
111: \newcommand\pic{{\rm Pic} \hskip1mm}
112: \newcommand\br{{\rm Br} \hskip1mm}
113: \newcommand \Z{{\mathbf Z}}
114: \newcommand\g{{\mathcal G}}
115: \newcommand\Div{{\rm Div}  }
116: \newcommand\X{{  \overline{X} }  }
117: \newcommand\Y{{  \overline{Y} }  }
118: \newcommand\U{{  \overline{U} }  }
119: \newcommand\et{\mathrm{\mathaccent 19 et}}
120: \newcommand\G{{\mathbf G}}
121: 
122: 
123: 
124: 
125: 
126: 
127: \begin{document}
128: 
129: 
130: \title[Rational points on cubic hypersurfaces]{Rational points on cubic hypersurfaces \\that
131:   split off a form}
132: 
133: 
134: 
135: 
136: 
137: \author{T.D. Browning}
138: \address{School of Mathematics\\
139: University of Bristol\\ Bristol BS8 1TW}
140: \email{t.d.browning@bristol.ac.uk}
141: 
142: 
143: \date{\today}
144: 
145: 
146: \begin{abstract}
147: Let $X$ be a projective  
148: cubic hypersurface of dimension $11$ or more,
149: which is defined over $\QQ$. We show that $X(\QQ)$ is non-empty provided
150: that the cubic form defining $X$ can be written as the sum of two 
151: forms that share no common variables.
152: \end{abstract}
153: 
154: \subjclass{11D72 (11E76, 11P55 14G25)}
155: 
156: 
157: \maketitle
158: \tableofcontents
159: 
160: 
161: 
162: 
163: 
164: \section{Introduction}
165: 
166: 
167: 
168: 
169: 
170: Let $X\subset\PP^{n-1}$ be a cubic
171: hypersurface, given as the zero locus of a cubic
172: form $C\in\ZZ[x_1,\ldots,x_n]$. This paper is concerned with the
173: problem of determining when the set of rational points $X(\QQ)$ on $X$ is
174: non-empty.  There is a well-known conjecture that $X(\QQ)\neq \emptyset$ as soon as
175: $n \geq 10$.  
176: In fact, for non-singular cubic hypersurfaces, it is expected that the
177: Hasse principle holds as soon as $n\geq 5$. This 
178: states that in order for $X(\QQ)$ to be non-empty it is necessary and
179: sufficient that $X(\QQ_p)$ is non-empty for every prime $p$. 
180: For a large class of possibly singular cubic hypersurfaces $X\subset \PP^{n-1}$, 
181: Salberger has calculated the Brauer group
182: $\mathrm{Br}(Y)$ associated to a projective non-singular model $Y$ of
183: $X$.  A detailed proof of this calculation is provided by 
184: Colliot-Th\'el\`ene in the appendix to this
185: paper. As a consequence of this investigation
186: one has the following prediction.
187: 
188: \begin{con}
189: Let $X\subset\PP^{n-1}$ be a cubic
190: hypersurface defined over $\QQ$ which is not a cone, with $n\geq 5$ and 
191: singular locus which is empty or of codimension at least $4$ in $X$.
192: %%IS THIS NECESSARY ASSUMPTION ON CONE?
193: Then the Hasse principle holds for the locus of non-singular points on $X$.
194: \end{con}
195: 
196: Let us now consider some of the progress that has been made towards
197: this conjecture. 
198: When $C$ is diagonal it follows from the work
199: of Baker \cite{baker} that $X$ has $\QQ$-rational points as soon as
200: $n\geq 7$. At the opposite end of the spectrum, when absolutely no assumptions are
201: made about the shape of $C$, a lot of work has been invested in producing a
202: reasonable lower bound for the number of variables needed to ensure that
203: $X(\QQ)\neq \emptyset$. Building on work of Davenport \cite{dav-32, dav-16},  Heath-Brown \cite{14} has 
204: recently shown that $n\geq 14$ variables are enough to
205: secure this property for an arbitrary cubic
206: hypersurface defined over $\QQ$.
207: In the light of this body of work it is very natural to try and evince
208: intermediate 
209: results in which the existence of rational points is guaranteed
210: for cubic hypersurfaces in fewer than $14$
211: variables when mild assumptions are made about the structure of the
212: hypersurface. It is precisely this point of view that is the focus of the present investigation. 
213: 
214: Let $\sing(X)$ denote the singular locus of $X$,
215: as a projective subvariety of $X$. When $C\in \ZZ[x_1,\ldots,x_n]$ is
216: non-singular, so that  $\sing(X)$ is empty, it follows from work of
217: Hooley \cite{hooley1} that the Hasse principle holds for $X$
218: provided that $n\geq 9$.
219: As is well-known, the local conditions are automatic when $n\geq 10$,
220: and so $X(\QQ)$ is non-empty for non-singular $X$ provided that $n\geq
221: 10$.  This fact was first proved by Heath-Brown \cite{hb-10}. 
222: When $\sing(X)$ has dimension
223: $\sigma\geq 0$, joint work of the author \cite{roth} with Heath-Brown 
224: shows that $X(\QQ)$ is non-empty provided that 
225: $$
226: n\geq 
227: \begin{cases}
228: 11, & \mbox{if $\sigma=0$,}\\
229: 12, & \mbox{if $\sigma=1$,}\\
230: 13, & \mbox{if $\sigma=2$.}
231: \end{cases}
232: $$
233: We will make use of this result shortly. 
234: 
235: 
236: Let $m<n$ be a positive integer. 
237: We will say that an integral cubic form $C$ in $n$ variables ``splits off an
238: $m$-form'' if there exist non-zero cubic forms $C_1,C_2$
239: with integer coefficients so that
240: $$
241: C(x_1,\ldots,x_n)=C_1(x_1,\ldots,x_m)+C_2(x_{m+1},\ldots,x_n),
242: $$
243: identically in $x_1,\ldots,x_n$. We will merely say that 
244: $C$ ``splits off a form'' if $C$ splits off an $m$-form for some
245: $1\leq m < n$.  The following is our 
246: main result. 
247: 
248: 
249: 
250: \begin{theorem}\label{main}
251: Let $X\subset \PP^{n-1}$ be a hypersurface defined by a cubic
252: form that splits off a form, with $n\geq 13$. Then $X(\QQ)\neq \emptyset$.
253: \end{theorem}
254: 
255: 
256: The essential content of Theorem \ref{main} is that we can save $1$
257: variable in the result of Heath-Brown \cite{14} when the
258: underlying cubic form splits off a form.
259: It should be stressed that 
260: the existence of a single $\QQ$-rational point on $X$ is
261: enough to demonstrate the $\QQ$-unirationality of $X$, 
262: and so the Zariski density of $X(\QQ)$ in $X$, when $X$ is
263: geometrically integral and not a
264: cone.   Variants of this result have been known for a long time (cf
265: \cite{ct2,manin, segre-b}). 
266: In the generality with which we have stated the result, it appears in 
267: the work of Colliot-Th\'el\`ene and Salberger
268: \cite[Proposition 1.3]{c-t-s} and in that of Koll\'ar \cite{kollar}. 
269: 
270: 
271: Our work has implications for the problem of
272: determining when an arbitrary cubic form $C\in \ZZ[x_1,\ldots,x_n]$ represents 
273: all non-zero $a\in \QQ$, using rational values for the variables. 
274: When this property holds we say that $C$
275: ``captures $\QQ^*$''.  Recall that a cubic form is said to be degenerate if 
276: the corresponding cubic hypersurface is a cone.
277: %%ie. there is a non-singular
278: %%linear transformation over $\cQ$ from $x_1,\ldots,x_n$ to $y_1,\ldots,y_n$ such that the transformed
279: %%form is one in $y_1,\ldots,y_{n-1}$ only.
280: %%that is, the coefficients of all terms containing
281: %%$y_n$ vanish. Degeneration is an absolute property, in the sense that if a form does
282: %%not degenerate by substitutions with coefficients in a particular field (e.g. the
283: %%rationals), this will remain true if the field is extended.
284: Fowler \cite{fowler} has shown that any
285: non-degenerate cubic
286: form that  represents $0$ automatically captures $\QQ^*$ provided only
287: that $n\geq 3$.  Hence it suffices to fix attention on those forms
288: that do not represent zero non-trivially. On multiplying through by
289: denominators it will clearly suffice to establish 
290: that cubic forms of the shape 
291: \begin{equation}
292:   \label{eq:1-form}
293: C(x_1,\ldots,x_n)-ax_{n+1}^3 
294: \end{equation}
295: represent zero non-trivially, with $a$ an
296: arbitrary non-zero integer.  But this form splits off a
297: $1$-form, and so is handled by  Theorem \ref{main}. In
298: this way we deduce the following result.
299: 
300: 
301: \begin{cor}
302: Let $C\in \ZZ[x_1,\ldots,x_n]$ be a non-degenerate cubic form, with
303: $n\geq 12$. 
304: Then $C$ captures $\QQ^*$.
305: \end{cor}
306: 
307: This result should be compared with the work of Heath-Brown \cite{14}, which
308: implies that $n\geq 13$ variables
309: suffice. It follows from the work of 
310: Hooley \cite{hooley1} that this may be improved to $n\geq 8$ 
311: when $C$ is non-singular. As indicated in \cite[Appendix 1]{hb-10} the
312: latter lower bound is 
313: probably the correct one for arbitrary cubic forms, 
314: since \eqref{eq:1-form} always has
315: non-trivial $p$-adic zeros for $n$ in this range.
316: 
317: 
318: Let $n\geq 4$. 
319: When $X\subset \PP^{n-1}$ is a hypersurface defined by a cubic
320: form that splits off a form, we are able to handle fewer 
321: variables when appropriate assumptions are made about one of the forms. 
322: If $X$ is a cone then we will see in Lemma \ref{lem:p1.1} that
323: $X(\QQ)\neq \emptyset$. If, on the other hand, $X$ is not a cone let
324: us consider the effect of supposing that 
325: the underlying cubic form splits off a
326: non-singular $m$-form $C_1(x_1,\ldots,x_m)$.
327: If $m=n-1$ then $X$ is itself non-singular and we automatically have
328: $X(\QQ)\neq \emptyset$ when $n\geq 10$. If $m\leq n-2$ then 
329: the residual form $C_2$ defines a projective cubic hypersurface of dimension $n-m-2$, and as such
330: has singular locus of dimension at most $n-m-3$.
331: But then it follows that $X$ has singular locus
332: of dimension at most $n-m-3$. Thus 
333: we may deduce from \cite{roth} that
334: $X(\QQ)\neq \emptyset$ provided that $n\geq 8+n-m$. 
335: We record this observation in the following result.
336: 
337: \begin{theorem}\label{t:goat}
338: Let $X\subset \PP^{n-1}$ be a hypersurface defined by a cubic
339: form that splits off a non-singular $m$-form, with $m\geq 8$ and
340: $n\geq 10$.
341: Then $X(\QQ)\neq \emptyset$.
342: \end{theorem}
343: 
344: 
345: It would be interesting to reduce the range of $m$ needed to ensure
346: the validity of Theorem \ref{t:goat}. 
347: Ours is not the first attempt to better understand the arithmetic of
348: cubic hypersurfaces that split off a form. 
349: Indeed, in Davenport's  \cite{dav-16} treatment of cubic forms in $16$ variables,
350: a fundamental ingredient in the treatment of certain
351: bilinear equations is a separate analysis of those forms that split
352: into two.  In further work, 
353: Colliot-Th\'el\`ene and Salberger \cite{c-t-s} have shown that the
354: Hasse principle holds for any cubic hypersurface in $\PP^{n-1}$ that contains a set of 
355: three conjugate singular points, provided only that $n\geq 4$.
356: Given a cubic extension $K$ of $\QQ$, define the corresponding norm form 
357: \begin{equation}
358:   \label{eq:NORM}
359: N(x_1,x_2,x_3):=\Norm_{K/\QQ}(\omega_1 x_1+\omega_2 x_2+\omega_3 x_3),
360: \end{equation}
361: where $\{\omega_1, \omega_2, \omega_3\}$ is a basis of $K$ as a vector
362: space over $\QQ$. In view of the fact that the local conditions are
363: automatically satisfied for cubic forms in at least $10$ variables, we
364: observe the following easy consequence.
365: 
366: 
367: 
368: \begin{theorem}[Colliot-Th\'el\`ene and Salberger 
369: \cite{c-t-s}]
370: \label{main'''}
371: Let $X\subset \PP^{n-1}$ be a hypersurface defined by a cubic
372: form that splits off a norm form, with $n\geq 10$. Then $X(\QQ)\neq \emptyset$.
373: \end{theorem}
374: 
375: 
376: 
377: 
378: %%As pointed out to the author by Professor Br\"udern, it seems likely
379: %%that one can handle as few as $n\geq 8$ variables when the cubic form
380: %%defining $X$ splits into the sum of a norm form and a non-singular
381: %%cubic form. 
382: It turns out that Theorem \ref{main'''} will play a
383: useful r\^ole in dispatching some of the cases that arise in the proof
384: of Theorem \ref{main}.   Following the strategy of 
385: Birch, Davenport and Lewis \cite{BDL}, 
386: it would however be straightforward to adapt the proof of Theorem~\ref{main}
387: to retrieve Theorem \ref{main'''}.
388: 
389: 
390: An obvious further line of enquiry would be to investigate cubic hypersurfaces that split
391: off two forms, by which we mean that the corresponding cubic form can be written
392: as 
393: $$
394: C(x_1,\ldots,x_n)=C_1(x_1,\ldots,x_\ell)+C_2(x_{\ell+1},\ldots,x_m)+C_3(x_{m+1},\ldots,x_n),
395: $$
396: identically in $x_1,\ldots,x_n$, for appropriate $1\leq \ell< m <n$.  With the extra structure apparent in
397: such hypersurfaces one would like to determine the most
398: general conditions possible under which the conjectured value of $n\geq 10$ variables suffices to
399: ensure the existence of $\QQ$-rational points. 
400: 
401: 
402: One of the remarkable features of our argument is
403: the breadth of tools that it draws upon. The underlying machinery is
404: the Hardy--Littlewood circle method,  and we certainly take advantage
405: of many of the contributions to the theory
406: of polynomial cubic exponential sums that have been made during the
407: last fifty years.  These are detailed in \S \ref{sec:exp_sums}.
408: A further component of our work 
409: involves a detailed analysis of the case in which one of the forms
410: that splits off in Theorem \ref{main} is singular and has a 
411: relatively small number of variables. To deal with this scenario 
412: it pays to reflect upon the classification of singular cubic hypersurfaces. This is a very old
413: topic in algebraic geometry, and can be traced back to the pioneering
414: work of Cayley \cite{cayley} and Schl\"afli \cite{cayley'}.
415: All of the necessary information will be collected together in \S \ref{s:geom}.
416: The final ingredient in  our work comprises good upper bounds for the
417: number of $\QQ$-rational points of bounded height on auxiliary cubic
418: hypersurfaces.  The estimates that we will take advantage of are
419: presented in \S \ref{sec:upper}. 
420: 
421: 
422: When it is applicable, the Hardy--Littlewood circle method allows
423: us to show that $X(\QQ)\neq\emptyset$ for a given cubic hypersurface $X\subset
424: \PP^{n-1}$ by evaluating asymptotically the number of $\QQ$-rational points of bounded
425: height on $X$. It is a well-known but intriguing feature of the method that 
426: one can achieve such precise information by first establishing weaker
427: upper bounds for the growth rate of $\QQ$-rational points on
428: appropriate auxiliary varieties. 
429: In fact, we will show in Lemma \ref{lem:8} that the 
430: $\QQ$-rational points on a non-singular cubic
431: hypersurface $X\subset \PP^{n-1}$ satisfy the growth bound
432: $$
433: \#\{x\in X(\QQ): H(x)\leq P\} \ll_{\ve, X} P^{\dim X-\frac{1}{2}+\ve},
434: $$
435: for any $P\geq 1$,  provided that $\dim X\geq 6$.  
436: Here $H: \PP^{n-1}(\QQ)\rightarrow \RR_{\geq 0}$ is the
437: usual exponential height function.  This should be compared with the
438: Manin conjecture \cite{f-m-t} which predicts that the 
439: exponent of $P$ should be  $\dim X-1$ as soon as $\dim X\geq 3$.
440: 
441: 
442: 
443: 
444: 
445: \begin{notat}
446: Throughout our work $\NN$ will denote the set of positive
447: integers.  For any $\al\in \RR$, we will follow common convention and
448: write $e(\al):=e^{2\pi i\al}$ and $e_q(\al):=e^{2\pi i\al/q}$. The
449: parameter $\ve$ will always denote a very small positive real
450: number.  We will use $|\x|$ to denote the norm $\max |x_i|$ 
451: of a vector $\x=(x_1,\ldots,x_n)\in \RR^n$, whereas $\|\x\|$ will
452: be reserved for the usual Euclidean norm $\sqrt{x_1^2+\cdots+x_n^2}$. 
453: All of the implied constants that appear in this 
454: work will be allowed to
455: depend upon the coefficients of the cubic forms under
456: consideration and the
457: parameter $\ve>0$.  Any further dependence will be explicitly
458: indicated by appropriate subscripts.
459: \end{notat}
460: 
461: 
462: 
463: 
464: 
465: \begin{ack}
466: It is a pleasure to thank 
467: Professor Colliot-Th\'el\`ene, Professor Heath-Brown, Professor Salberger and 
468: Professor Wooley for several useful discussions.  
469: While working on this paper the author was supported by EPSRC
470: grant number \texttt{EP/E053262/1}.
471: \end{ack}
472: 
473: 
474: 
475: \section{Geometry of singular cubic hypersurfaces}\label{s:geom}
476: 
477: 
478: 
479: The proof of Theorem \ref{main} will depend intimately on the  
480: dimension of the hypersurfaces defined by the constituent cubic
481: forms, and the nature of their singularities. 
482: A key step will be to determine conditions on this
483: singular locus under which the hypersurface 
484: automatically has rational points.
485: 
486: Let $C\in \ZZ[x_1,\ldots,x_n]$ be an arbitrary cubic form, which
487: we assume takes the shape
488: \begin{equation}
489:   \label{eq:C}
490: C(\x):=\sum_{i,j,k}c_{ijk}x_ix_jx_k,
491: \end{equation}
492: in which the coefficients $c_{ijk}\in \ZZ$ are symmetric in the indices $i,j,k$.
493: Define the $n\times n$ matrix $M(\x)$ with $j,k$-entry
494: $\sum_i c_{ijk}x_i.$  We will say that the cubic form $C$ is ``good'' if 
495: for any $H\geq 1$ and any $\ve>0$ we have the upper bound
496: $$
497: \#\{\x\in\ZZ^{n}: |\x|\leq H,\,\rank M(\x)=r\}\ll H^{r+\ve},
498: $$
499: for each integer $0\leq r\leq n$.  A crucial step in Davenport's \cite{dav-16}
500: treatment of general cubic forms is a proof of the fact that forms
501: that fail to be good automatically possess non-trivial integer
502: solutions for ``geometric reasons''.  
503: Our approach has a similar flavour, although the underlying arguments will
504: be more obviously geometric.
505: 
506: 
507: 
508: Assume throughout this section that $n \geq 3$ and $X\subset
509: \PP^{n-1}$ is a hypersurface defined by a cubic form $C\in \ZZ[x_1,\ldots,x_n]$.
510: A lot of the facts
511: that we will record are classical. 
512: Suppose for the moment that $C$ is not absolutely irreducible. Then
513: either it has a linear factor $L$ defined over $\QQ$, or it is a product
514: $C=L_1L_2L_3$ of three linear factors that are conjugate over $\cQ$.
515: By considering the equation $L=0$ in the former case, 
516: we deduce that $X(\QQ)\neq\emptyset$. 
517: In the latter case, we arrive at the same conclusion when $n\geq 4$, by
518: considering the system of equations $L_1=L_2=L_3=0$. 
519: When $n=3$ and $C$ is a product of three conjugate factors we deduce
520: that $X$ has precisely three conjugate singular points.  
521: When $n\geq 3$ and $X$ is defined by an absolutely
522: irreducible cubic form, but is a cone, we note that the space of
523: vertices on $X$ must be a linear space globally defined over $\QQ$.
524: Thus $X(\QQ)\neq \emptyset$ in this case too. 
525: We have therefore established the following simple result.
526: 
527: 
528: \begin{lemma}\label{lem:p1.1}
529: Let $n \geq 4$. If $X$ is not geometrically integral, or if
530: $X$ is a cone, then $X(\QQ)\neq \emptyset$. When $n=3$ the same
531: conclusion holds unless $X$ contains precisely three conjugate singular points.
532: \end{lemma}
533: 
534: 
535: Recall that a cubic hypersurface $X$ is said to 
536: be non-singular if over $\cQ^n$ the only solution to the
537: system of equations $\nabla C(\x)=\mathbf{0}$ has $\x=\mathbf{0}$. 
538: Henceforth we will be predominantly interested in 
539: singular cubic forms, and then only in the cases $n=3,4$ and $5$.
540: Let $k$ be a field. It has been conjectured by 
541: Cassels and Swinnerton-Dyer that any cubic hypersurface $X \subset \PP^{n-1}$ defined over $k$ that contains a
542: $k$-rational $0$-cycle of degree coprime to $3$, automatically has
543: a $k$-rational point.  The case $n=3$ goes back to Poincar\'e. When
544: the singular locus is non-empty, the case $n=4$  can be 
545: deduced from the work of Skolem \cite{skolem}. A comprehensive
546: discussion of the arithmetic of singular cubic surfaces can be found
547: in the work of Coray and Tsfasman \cite{c-t}. 
548: Coray \cite{coray76b} has established the conjecture for all local
549: fields and,  in a subsequent investigation \cite[Proposition 3.6]{coray87}, has also dispatched
550: the case in which  $n=5$ and the $0$-cycle is made up of double points. 
551: 
552: 
553: Suppose first that $n=3$, so that $X\subset \PP^2$ defines a cubic
554: curve, which we assume to be geometrically integral and not a cone. 
555: When $X$ is singular it contains exactly one 
556: singular point, which must therefore be defined over $\QQ$. Once
557: combined with Lemma \ref{lem:p1.1} we arrive at the following result.
558: 
559: 
560: \begin{lemma}\label{lem:curve}
561: Let $n=3$ and suppose that $X(\QQ)= \emptyset$. Then 
562: \begin{enumerate}
563: \item[(i)]
564: $X$ is non-singular; or
565: \item[(ii)]
566: $X$ contains precisely three conjugate singular points. 
567: \end{enumerate}
568: \end{lemma}
569: 
570: In case (ii) of Lemma \ref{lem:curve} one concludes that
571: the underlying cubic form can be written as a norm form \eqref{eq:NORM},
572: for appropriate $\omega_1,\omega_2,\omega_3 \in K$, 
573: where $K$ is the cubic number field obtained by adjoining one
574: of the singularities. 
575: 
576: 
577: We turn now to the case $n=4$ of cubic surfaces 
578: $X\subset \PP^3$, which we suppose to be geometrically integral and
579: not equal to a cone. Suppose that $X$ is singular. The classification
580: of such cubic surfaces can be traced back to Cayley \cite{cayley} and Schl\"afli \cite{cayley'}, but we
581: will employ the modern treatment found in the work of Bruce and
582: Wall \cite{b-w}. In particular the singular locus of $X$ is either a
583: single line, in which case $X$ is ruled, or else it contains $\delta\leq 4$
584: isolated singularities and these are all rational double points. 
585: It follows that $X(\QQ)\neq \emptyset$ unless $\delta=3$ and the
586: three singular points are conjugate to each other over a cubic extension of $\QQ$.
587: In this final case, Skolem \cite{skolem} showed that
588: $C$ can be written as
589: \begin{equation}
590:   \label{eq:4_norm}
591: \Norm_{K/\QQ}(x_1\omega_1+x_2\omega_2+x_3\omega_3)+ax_4^2\Tr_{K/\QQ}(x_1\omega_1+x_2\omega_2+x_3\omega_3)+bx_4^3,
592: \end{equation}
593: for appropriate coefficients $\omega_1,\omega_2,\omega_3 \in K$ and
594: $a,b \in \ZZ$, where $K$ is the cubic number field obtained by adjoining one
595: of the singularities to $\QQ$. In terms of the classification over $\cQ$ 
596: according to singularity type, the 
597: only possibility here is that $X$ has
598: singularity type $3\mathbf{A}_i$ for $i=1$ or $2$, since the action of
599: $\Gal(\cQ/\QQ)$ preserves the singularity type.
600: Bringing this all together, we have therefore established the
601: following analogue of Lemma \ref{lem:curve}.
602: 
603: 
604: 
605: \begin{lemma}\label{lem:surface}
606: Let $n=4$ and suppose that $X(\QQ)= \emptyset$. Then 
607: \begin{enumerate}
608: \item[(i)] 
609: $X$ is non-singular; or
610: \item[(ii)]
611: $X$ contains precisely three conjugate double points. 
612: \end{enumerate}
613: \end{lemma}
614: 
615: 
616: 
617: 
618: We now try to construct a version of Lemmas \ref{lem:curve} and
619: \ref{lem:surface} for the case $n=5$. 
620: Let $Y\subset X$ denote the singular locus of $X\subset \PP^4$, a variety of
621: dimension at most $2$.  As usual we assume that $X$ is geometrically
622: integral and not a cone.  
623: We analyse $Y$ by considering the intersection of $X$ with a generic
624: hyperplane $H\in {\PP^4}^*$.  
625: In particular the hyperplane section 
626: $$
627: S_H=H \cap X
628: $$ 
629: is a geometrically integral cubic surface which is not a cone (see
630: \cite[Proposition 18.10]{harris}, for example).
631: In taking $H$ to be defined over $\QQ$, we may further assume that
632: $S_H$ is defined over $\QQ$. Any $\QQ$-rational point on $S_H$ visibly
633: produces a $\QQ$-rational point on $X$. 
634: Let $T_H  \subset S_H$ denote the singular locus of $S_H$. 
635: Then the classification of cubic surfaces 
636: implies that $T_H$ is either empty or it is a union of $\delta_H\leq
637: 4$ points or it is a line. 
638: When $T_H$ is non-empty it follows from Lemma~\ref{lem:surface}
639: that either $S_H(\QQ)\neq \emptyset$ or else $T_H$ is finite, with 
640: $\delta_H=3$ and the three points being conjugate
641: to each other over $\cQ$.
642: 
643: Now an application of Bertini's theorem (in the form given by Harris 
644: \cite[Theorem~$17.16$]{harris}, for example) shows that
645: $$
646: H\cap Y =T_H.
647: $$
648: When $S_H$ is non-singular it therefore follows that $H\cap Y$ is
649: empty for generic   $H\in {\PP^4}^*$, whence $Y$ must be finite. In
650: the alternative case, when
651: $S_H$ is singular, we may conclude that $\#(H\cap Y)=3$ for 
652: generic   $H\in {\PP^4}^*$, whence the maximal component of $Y$
653: is a cubic curve.
654: 
655: Let us examine further the possibility that the singular locus $Y$ of
656: $X$ has dimension $1$, and that it contains a cubic curve $Y_0$ as its
657: component of maximal dimension. Clearly $Y_0$ is defined over $\QQ$.
658: Furthermore,  we may conclude from B\'ezout's theorem that the line connecting
659: any two points of $Y_0$ must be contained in $X$, since each such
660: point is a singularity of $X$. 
661: 
662: 
663: If $Y_0$ is reducible over $\QQ$ then there are two
664: basic possibilities: either it is a union of lines or it is a union of
665: a conic and a line.  In the latter case $Y_0$ contains a line defined
666: over $\QQ$ and it trivially follows
667: that $Y_0(\QQ)\neq \emptyset$.  The former case fragments into a number
668: of subcases: either it is a union of $3$ concurrent lines,  or it contains
669: a pair of skew lines, or it is a union of $3$ coplanar lines, or it
670: contains a repeated line. The second case is 
671: impossible since then the join of the two skew lines defines a
672: $3$-plane  that would also be contained in $X$, contradicting the fact that $X$ is geometrically
673: irreducible.  It follows from consideration of the Galois
674: action on $Y_0$ that $Y_0(\QQ)\neq \emptyset$ in every case apart from
675: the one in which $Y_0$ is a union of $3$ coplanar lines. 
676: 
677: 
678: If $Y_0$ is geometrically irreducible then it cannot be a twisted
679: cubic since then the secant variety $S(Y_0)\cong
680: \PP^3$ would be contained in $X$. 
681: Our argument so far has shown that either $Y_0(\QQ)\neq \emptyset$ for trivial
682: reasons, or else $Y_0$ is a cubic plane curve that is either geometrically
683: irreducible or a union of $3$ distinct lines.
684: The plane $P$ containing $Y_0$ is defined over $\QQ$ and, after carrying out a
685: linear change of variables, we may take
686: $x_1=x_2=0$ as its defining equations. But then it follows
687: that the cubic form defining $X$ can be written  
688: $$
689: x_1Q_1(x_1,\ldots, x_5)+x_2Q_2(x_1,\ldots, x_5),
690: $$
691: for appropriate quadratic forms $Q_1,Q_2$ defined over $\ZZ$. 
692: With this notation one sees that $Y$ is the locus of
693: solutions to the system of equations 
694: $$
695: x_1=x_2=Q_1(0,0,x_3,x_4,x_5)=Q_2(0,0,x_3,x_4,x_5)=0,
696: $$
697: in $\PP^4$. It is now clear that 
698: the component $Y_0$ of $Y$ of maximal
699: dimension cannot be a cubic plane curve of the two remaining types.
700: 
701: 
702: It remains to deal with the case in which the singular locus $Y$ of $X$
703: is finite and globally defined over $\QQ$. As shown by C. Segre
704: \cite{segre}, we have $\delta=\#Y\leq 10$, the extremal
705: case of $10$ singular points being achieved by the so-called
706: Segre threefold.  Since $X$ is assumed not to be a cone so we may
707: assume that all of the singularities are double points. Indeed any
708: singularity with multiplicity exceeding $2$ must be a vertex for
709: $X$. In fact, when $\delta\geq 6$ it is known \cite[Lemma 2.2]{6} that
710: all the singularities are actually nodal.  
711: Appealing to Coray's partial
712: resolution of the Cassels--Swinnerton-Dyer conjecture for threefolds,
713: we are now ready to record our analogue of Lemmas \ref{lem:curve} and \ref{lem:surface}.
714: 
715: \begin{lemma}\label{lem:3fold}
716: Let $n=5$ and suppose that $X(\QQ)= \emptyset$. Then 
717: \begin{enumerate}
718: \item[(i)]
719: $X$ is non-singular; or
720: \item[(ii)]
721: $X$ is a geometrically integral cubic
722: hypersurface whose singular locus contains precisely $\delta$ double
723: points, with 
724: $\delta\in \{3,6, 9\}$.
725: \end{enumerate}
726: \end{lemma}
727: 
728: 
729: 
730: In the second case of Lemma \ref{lem:3fold} it follows from
731: \cite{c-t-s} and \cite{6} that the Hasse principle holds for $X$ when
732: $\delta=3$ or $6$. Our investigation would be made easier if we were
733: also in possession of this fact when $\delta=9$.
734: Lacking this, all that we actually require from part (ii) of Lemma
735: \ref{lem:3fold} is that the singular locus should be finite. 
736: In his survey of open problems in Diophantine geometry, Lewis \cite{lewis}
737: reports on unpublished joint work with Blass, which would appear to give
738: Lemma \ref{lem:3fold}. However, in the absence of subsequent
739: elucidation, 
740: we have chosen to present our own proof of this result.
741: 
742: 
743: \section{Cubic exponential sums}\label{sec:exp_sums}
744: 
745: 
746: Let $C\in \ZZ[x_1,\ldots,x_n]$ be an arbitrary
747: cubic form, assumed to take the shape \eqref{eq:C}. 
748: Our work in this section centres upon various properties of the cubic
749: exponential sums
750: \begin{equation}\label{eq:Tcubic}
751: S(\al)=S_w(\al;C,P):=\sum_{\x\in\ZZ^n}w(P^{-1}\x)e(\al C(\x)),
752: \end{equation}
753: for a suitable family of weights $w$ on $\RR^n$, and cubic forms 
754: that are always either good (in the sense of the previous section) or the 
755: hypersurface they define has finite (possibly empty) singular locus.
756: Specifically, we will 
757:  collect together some general upper bounds for $S(\al)$, some estimates for suitable moments of $S(\al)$
758: and some asymptotic formulae for $S(\al)$ when suitable assumptions
759: are made about how $\al$ can be approximated by rational numbers.
760: All of these estimates will depend on the parameter $P$ which should
761: be thought of as tending to infinity.
762: 
763: We must begin by saying a few words about the weight functions that we
764: will be working with. 
765: Let $n_1,n_2\geq 0$ such that $n_1+n_2=n$. When $n_i\geq 1$ we let $\z_i\in\RR^{n_i}$
766: be certain vectors, which we think of as being fixed, but whose
767: nature will be determined later. Similarly we let $\rho>0$. 
768: All of the estimates in our work will be allowed to depend upon the
769: choice of $\z_1,\z_2$ and $\rho$. Define $w_1:\RR^{n_1}\rightarrow
770: \RR_{\geq 0}$, via 
771: \begin{equation}
772:   \label{eq:w1}
773: w_1(\x_1):=\exp(-\|\x_1-\z_1\|^2(\log P)^4),
774: \end{equation}
775: where we have written $\x_1=(x_1,\ldots,x_{n_1})$.
776: Let $P_0=P(\log P)^{-2}$. 
777: Then 
778: $$
779: w_1(P^{-1}\x_1)=
780: \exp(-\|\x_1-P\z_1\|^2 P_0^{-2})
781: $$
782: is exactly the weight function introduced by Heath-Brown
783: in \cite[\S 3]{hb-10}.  
784: Note that 
785: $$
786: \nabla w_1(\x_1)=-2(\log P)^4 w_1(\x_1)(x_1-z_1,\ldots,x_{n_1}-z_{n_1}),
787: $$
788: so that  $\nabla w_1(\x_1)\ll (\log P)^4$ for
789: any $\x_1\in\RR^{n_1}$.
790: Next we let $w_2:\RR^{n_2}\rightarrow \{0,1\}$ denote the 
791: characteristic function 
792: \begin{equation}
793:   \label{eq:w2}
794:   w_2(\x_2):=
795: \begin{cases}
796: 1, &\mbox{if $|\x_2-\z_2|<\rho$,}\\
797: 0, &\mbox{otherwise,}
798: \end{cases}
799: \end{equation}
800: where
801: $\x_2=(x_{n_1+1},\ldots,x_{n})$.
802: 
803: Each weight $w$ appearing in our work will either be of the
804: form $w_1$ or $w_2$ or $w=(w_1,w_2)$, depending on context. To help
805: distinguish which estimates are valid for which choice of weight
806: function, let us denote by $\WW_{n}^{(1)}$ the set of non-negative weight functions on
807: $\RR^n$  that are of the shape \eqref{eq:w1}, and let 
808: $\WW_n^{(2)}$ denote the corresponding set of weight functions on
809: $\RR^n$  of the type \eqref{eq:w2}. We let $\WW_n$ denote the set of
810: mixed functions $w=(w_1,w_2)$, with $w_i\in \WW_{n_i}^{(i)}$ for $i=1,2$.
811: In particular $\WW_{n}^{(i)}\subset \WW_n$ for $i=1,2$.
812: In the definition of these sets the precise value of $\z_1,\z_2$ or $\rho$
813: is immaterial, unless explicitly indicated otherwise, and the
814: corresponding implied constants will always be allowed to depend on these
815: quantities in any way. 
816: 
817: We are now ready to record the upper bounds for $S(\al)$ that feature
818: in our investigation. 
819: 
820: \begin{lemma}\label{lem:B}
821: Let $\ve>0$, let $w\in \WW_n$ and assume that $C\in \ZZ[x_1,\ldots,x_n]$ is a good cubic form.
822: Let $a,q\in \ZZ$ such that $0\leq a< q\leq P^{3/2}$ and
823: $\gcd(a,q)=1$. Then if $\al=a/q+\theta$ we have
824: $$
825: S(\al)\ll P^{n+\ve}\big( q|\theta| +(q|\theta|P^3)^{-1}\big)^{\frac{n}{8}}.
826: $$
827: If furthermore $|\theta|\leq q^{-1}P^{-3/2}$, then we have 
828: $$
829: S(\al)\ll P^{n+\ve}q^{-\frac{n}{8}}\min\{1,(|\theta|P^3)^{-\frac{n}{8}}\}.
830: $$
831: \end{lemma}
832: 
833: \begin{proof}
834: This is the essential content of the investigation of Davenport
835: \cite{dav-16} into cubic forms in $16$ variables. 
836: The bounds are derived in a more succinct manner in
837: Heath-Brown \cite[\S 2]{14}. The fact that we are
838: working with exponential sums that are differently weighted
839: makes no difference to the validity of the argument, and the
840: reader may wish to consult \cite[\S 9]{41}, where the necessary
841: modifications can be found in the setting of quartic forms. 
842: \end{proof}
843: 
844: 
845: Define the complete exponential sum 
846: \begin{equation}
847:   \label{eq:complete}
848: S_{a,q}:=\sum_{\y \bmod{q}}e_q(aC(\y)),
849: \end{equation}
850: for any coprime integers $a,q$ such that $q>0$. It can easily  be
851: deduced from the proof of Lemma \ref{lem:B} that $S_{a,q}\ll
852: q^{7n/8+\ve}$ for any $\ve>0$, under the assumption that the cubic
853: form is good. The following improvement is due to Heath-Brown \cite[\S 7]{14}.
854: 
855: 
856: \begin{lemma}\label{lem:paris}
857: Let $\ve>0$  and assume that $C\in \ZZ[x_1,\ldots,x_n]$ is a good
858: cubic form. Then we have $S_{a,q}\ll q^{5n/6+\ve}$.
859: \end{lemma}
860: 
861: 
862: We now come to the real workhorse 
863: in our argument.  Given $R,\phi>0$ and $v>0$ we define
864: \begin{equation}\label{eq:sig}
865: \Sigma_v(R,\phi,\pm):=  
866: \sum_{R<q\leq 2R} \sum_{\substack{a\bmod{q}\\
867:     \gcd(a,q)=1}}\int_{\phi}^{2\phi} \big|S\big(\frac{a}{q}\pm t\big)\big|^v \d t.
868: \end{equation}
869: The following result provides an upper bound for this quantity.
870: 
871: \begin{lemma}\label{lem:D}
872: Let $\ve>0$, let $w\in \WW_n$ 
873:  and assume that $C\in \ZZ[x_1,\ldots,x_n]$ is a good cubic form.
874: Let $R,\phi>0$, with $R\leq P^{3/2}$ and $\phi\leq R^{-2}$.
875: Then for any $v\in [0,2]$ and any $H\in
876: [1,P]\cap \ZZ$ we have
877: \begin{align*}
878: \Sigma_v(R,\phi,\pm)\ll P^3
879: +R^2\phi^{1-\frac{v}{2}} \left(\frac{\psi_H P^{2n-1+\ve}}{H^{n-1}}F\right)^{\frac{v}{2}},
880: \end{align*}
881: where
882: $$
883: \psi_H:=\phi+\frac{1}{P^2H},
884: \quad
885: F:=1+(RH^3\psi_H)^{\frac{n}{2}}+\frac{H^n}{R^{\frac{n}{2}}(P^2\psi_H)^{\frac{n-2}{2}}}.
886: $$
887: \end{lemma}
888: 
889: \begin{proof}
890: It is clear that $\Sigma_0(R,\phi,\pm)\ll R^2\phi.$
891: Hence it follows from H\"older's inequality that
892: \begin{align*}
893: \Sigma_v(R,\phi,\pm)&\ll
894: (R^2\phi)^{1-\frac{v}{2}} \Sigma_2(R,\phi,\pm)^{\frac{v}{2}}.
895: \end{align*}
896: On employing Heath-Brown's estimate for
897: $\Sigma_2(R,\phi,\pm)$, 
898: which follows from \cite[Eqs. (4.5) and (5.1)]{14},
899: we therefore deduce that
900: \begin{align*}
901: \Sigma_v(R,\phi,\pm)\ll
902: (R^2\phi)^{1-\frac{v}{2}} \left(\psi_H R^2\Big( P^2H + \frac{P^{2n-1+\ve}}{H^{n-1}}F\Big)\right)^{\frac{v}{2}}.
903: \end{align*}
904: As in the deduction of Lemma \ref{lem:B}, the fact that we are
905: working with differently weighted exponential sums makes no
906: difference to the final outcome of the argument.
907: 
908: Using the fact that $R\leq P^{3/2}$ and  $\phi\leq R^{-2}$, with $H\leq P$,
909: it easily follows that the term involving $P^2H$ contributes
910: \begin{align*}
911: \ll 
912: (R^2\phi)^{1-\frac{v}{2}}(\phi R^2 P^3 +R^2)^{\frac{v}{2}}\ll 
913: R^2\phi P^{\frac{3v}{2}}+ R^2\phi^{1-\frac{v}{2}}\ll P^3,
914: \end{align*}
915: since $0\leq v\leq 2$. This completes the proof of the lemma. 
916: \end{proof}
917: 
918: 
919: 
920: Lemma \ref{lem:D} is based on an averaged version of van der Corput's
921: method and comprises the key innovation in the work of Heath-Brown
922: \cite{14} already alluded to. Although we have presented it in the
923: context of denominators $q$ and values of $\al=a/q\pm t$ restricted to
924: dyadic intervals, the general result consists of a bound for 
925: $\int |S(\al)|^2 \d\al$, where the integral is taken over a certain
926: set of minor arcs. For cubic forms in few variables
927: we will have better results available. When $n=1$ and $w\in \WW_1^{(2)}$, Hua's inequality \cite[Lemma 3.2]{dav-book} implies that
928: $$
929: \int_{0}^1|S(\al)|^{2^j}\d\al \ll P^{2^j-j+\ve},
930: $$
931: for any $j\leq 3$ . The following result is due to 
932: Wooley \cite{wooley}, and generalises this to binary forms.
933: 
934: \begin{lemma}\label{lem:hua_2}
935: Let $\ve>0$, let $w\in\WW_2^{(2)}$ and let 
936: $C\in \ZZ[x_1,x_2]$ be a binary cubic form, not of the shape
937: $a(b_1x_1+b_2x_2)^3$, for integers $a,b_1,b_2$. 
938: Then we have 
939: $$
940: \int_{0}^1|S(\al)|^{2^{j-1}}\d\al \ll P^{2^j-j+\ve},
941: $$
942: for any $j\leq 3$.
943: \end{lemma}
944: 
945: 
946: Our next selection of results concern the approximation of $S(\al)$ on a certain set of arcs in
947: the interval $[0,1]$.  For given $A,B,C\geq 0$, define 
948: $\mathcal{A}=\mathcal{A}(A,B,C)$ to be set of $\al\in [0,1]$ for
949: which there exists $a,q\in \ZZ$ such that $0\leq a< q\leq P^A$ and
950: $\gcd(a,q)=1$, with 
951: \begin{equation}
952:   \label{eq:Aqa}
953: \al\in \A_{q,a}:=\Big[\frac{a}{q}-\frac{1}{q^{B}P^{3-C}},  
954: \frac{a}{q}+\frac{1}{q^{B}P^{3-C}}\Big].
955: \end{equation}
956: The major arcs in our work will be a subset of these, but it will be
957: useful to maintain a certain degree of generality.
958: When dealing with cubic forms whose singular locus is very small, we
959: have rather good control over the approximation of $S(\al)$ on the
960: arcs $\A=\A(A,B,C)$, provided that we work with the
961: class of smooth weight functions $\WW_n^{(1)}$.  
962: Recall the definition of $S_{a,q}$ from \eqref{eq:complete} and let
963: $$
964:  I_w(\psi):=\int_{\RR^n}w(\x)e(\psi C(\x))\d\x,
965: $$
966: for $\psi\in\RR$.
967: We will need to work with the familiar quantity
968: \begin{equation}
969:   \label{eq:S*}
970: S^*(\al):=q^{-n}P^n S_{a,q}I_w(\theta P^3),
971: \end{equation}
972: concerning which we have the following result.
973: 
974: 
975: \begin{lemma}\label{lem:F}
976: Let $\ve>0$ and $n\geq 3$. Assume that
977: $C\in\ZZ[x_1,\ldots,x_n]$ is a good
978: cubic form defining a projective hypersurface that is not a cone, with singular locus of
979: dimension $\sigma\in \{-1,0\}$.
980: Let $A,B,C\geq 0$ such that $A<1$ and $B\in \{0,1\}$, and let $\al\in
981: \A_{q,a}$.  Then there exists $w\in\WW_n^{(1)}$  such that 
982: $$
983: S(\al)-S^*(\al)\ll P^{A(\frac{n}{3}+\frac{\sigma}{2})+\frac{n+1}{2}+\ve}
984: +P^{A(1-B)\frac{n+1+\sigma}{2}+C\frac{n+1}{2}+\ve}.
985: $$
986: Furthermore, if $kn\geq 12$ and $k\leq 9$, then we have 
987: $$
988: \int_\A |S^*(\al)|^k \d\al\ll P^{kn-3+\ve}.
989: $$
990: \end{lemma}
991: 
992: \begin{proof}
993: The proof of this result is based on the investigation carried out by
994: Heath-Brown \cite{hb-10} into non-singular cubic forms in $10$ variables.
995: One of the key ingredients in his approach is the Poisson summation
996: formula, and it is this part of the argument that we plan to take
997: advantage of. 
998: 
999: We begin by choosing $\z_1\in \RR^{n}$ to be a point at which the
1000: matrix of second derivatives of $C$ has full rank 
1001: at $\z_1$.  The existence of such a point follows from
1002: the work of Hooley \cite[Lemma 26]{hooley2}. 
1003: With this choice of $\z_1$ we now select $w$ to be the weight function
1004: in \eqref{eq:w1}, which belongs to $\WW_n^{(1)}$. Let 
1005: \begin{align*}
1006: S_{a,q}(\v)&:=\sum_{\y \bmod{q}}e_q(aC(\y)+\v.\y),\\
1007: J_w(\psi,\v)&:=\int_{\RR^n}w(P^{-1}\x)e(\psi C(\x)-\v.\x)\d\x,
1008: \end{align*}
1009: for any $\v\in \RR^n$, and let $\al=a/q+\theta \in \A_{q,a}$.
1010: Then \cite[Lemma 8]{hb-10} yields
1011: $$
1012: S(\al)-S^*(\al)\ll 1+ q^{-n}\sum_{\substack{\v\in \ZZ^n\\1\leq |\v|\ll
1013:   V}} S_{a,q}(\v)J_w(\theta,q^{-1}\v),
1014: $$
1015: where
1016: $
1017: V:=(\log P)^7 q (P^{-1}+|\theta|P^2),
1018: $
1019: and furthermore,
1020: \begin{equation}
1021:   \label{eq:bound_99}
1022: J_w(\theta,\mathbf{w})\ll P^n(\log P)^{7n}\min \{1, (|\theta|P^3)^{-1}\}^{\frac{n-1}{2}},
1023: \end{equation}
1024: for any $\mathbf{w}\in \RR^n$.
1025: The main  difference between what we have recorded here and the
1026: statement of  \cite[Lemma 8]{hb-10} is that our definition of $S(\al)$
1027: does not involve a summation over $a$. This
1028: deviation makes no difference to the final outcome.
1029: Note that once the existence of a suitable point $\z_1$ is established
1030: for the definition of the weight function, the manipulations  involving
1031: the exponential integral remain valid even when $C$ is singular. 
1032: 
1033: The summation over $\v$ in our upper bound for $S(\al)-S^*(\al)$
1034: implies in particular $V\gg 1$. Since $A<1$ we automatically have 
1035: $(\log P)^7 q P^{-1}\leq (\log P)^7P^{A-1}=o(1)$. 
1036:  Hence the condition $V\gg 1$ implies that
1037: $$
1038: (\log P)^7 q |\theta|P^2\leq  V\ll (\log P)^7 q |\theta|P^2
1039: $$
1040: and 
1041: \begin{equation}
1042:   \label{eq:size}
1043: q^{-1}P^{-2}(\log P)^{-7}\ll |\theta| \leq q^{-B}P^{-3+C}.  
1044: \end{equation}
1045: Putting everything together it follows that
1046: $$
1047: S(\al)-S^*(\al)\ll 1+ q^{-n}(\log
1048: P)^{7n}P^{\frac{3-n}{2}}|\theta|^{\frac{1-n}{2}}\mathcal{T}(V),
1049: $$
1050: where
1051: $$
1052: \mathcal{T}(V):=\sum_{1 \leq |\v|\ll V} |S_{a,q}(\v)|.
1053: $$
1054: We will show that
1055: \begin{equation}\label{eq:JB}
1056: \mathcal{T}(V)\ll q^{\frac{n+1+\sigma+\ve}{2}}(V^n+ q^{\frac{n}{3}}),
1057: \end{equation}
1058: for $\sigma\in \{-1,0\}$. 
1059: Before doing so let us see how this suffices to complete the proof of
1060: the first part of the lemma.  Recalling from above that $V$ has
1061: order of magnitude $(\log P)^7q|\theta|P^2$, 
1062: and employing \eqref{eq:size}, we deduce that
1063: \begin{align*}
1064: S(\al)-S^*(\al)
1065: \ll& q^{-\frac{n}{2}+\frac{1+\sigma}{2}+\frac{2\ve}{3}} P^{-\frac{n-3}{2}}|\theta|^{-\frac{n-1}{2}}
1066: ((q|\theta|P^2)^{n}+ q^{\frac{n}{3}})\\
1067: \ll& q^{\frac{n+1+\sigma}{2}+\frac{2\ve}{3}}|\theta|^{\frac{n+1}{2}}P^{\frac{3(n+1)}{2}}+
1068: q^{-\frac{n}{6}+\frac{1+\sigma}{2}+\frac{2\ve}{3}}|\theta|^{-\frac{n-1}{2}}P^{-\frac{n-3}{2}}\\
1069: \ll& q^{(1-B)\frac{n}{2}-\frac{B}{2}+\frac{1+\sigma}{2}}P^{C\frac{n+1}{2}+\ve}
1070: +P^{A(\frac{n}{3}+\frac{\sigma}{2})+\frac{n+1}{2}+\ve}.
1071: \end{align*}
1072: If $B=1$ then the first term here is $O(P^{C\frac{n+1}{2}+\ve})$,
1073: since $\sigma\leq 0$.
1074: Alternatively, if $B=0$, then the first term is 
1075: $
1076: O(P^{A\frac{n+1+\sigma}{2}+C\frac{n+1}{2}+\ve}).
1077: $
1078: This establishes the first part of the lemma subject to \eqref{eq:JB}.
1079: 
1080: To establish \eqref{eq:JB} we return to the manipulations in
1081: \cite[\S 5]{41}. Things are simplified slightly by no longer needing
1082: to keep track of the dependence on $C$ in each implied
1083: constant.  In particular we may take $H\ll 1$ throughout.
1084: The sum $S_{a,q}(\v)$ satisfies a basic multiplicativity property, as
1085: recorded in \cite[Lemma~10]{41}. 
1086: Write $q=bc^2d$, where 
1087: $$
1088: b:=\prod_{\substack{p^e\| q\\ e\leq 2}} p^e, \quad
1089: d:=\prod_{\substack{p^e\| q\\ e\geq 3, 2\nmid e}} p.
1090: $$
1091: In particular $d\mid c$ and we deduce from \cite[Lemmas~7, 10 and 11]{41} that
1092: \begin{align*}
1093: \mathcal{T}(V)
1094: &\ll q^{\frac{n}{2}}b^{\frac{1+\sigma}{2}+\frac{\ve}{2}}\sum_{1 \leq |\v|\ll V}
1095: \sum_{\substack{\mathbf{a}\bmod c\\ c\mid (a\nabla C(\mathbf{a})+\v)}}
1096: N_d(\mathbf{a})^{\frac{1}{2}}.
1097: \end{align*}
1098: Here, if $M(\x)$ denotes the matrix of second derivatives of $C(\x)$,
1099: then $N_m(\mathbf{x})$ is the number of $\y$ modulo $m$ such that
1100: $M(\mathbf{x})\y \equiv \mathbf{0} \bmod{m}$.
1101: Recalling the notation of \cite[Lemma 12]{41},
1102: in which we take $\v_0=\mathbf{0}$ and $g=C$, 
1103: it follows that there is
1104: an absolute constant $\kappa>0$ such that 
1105: \begin{align*}
1106: \mathcal{T}(V)
1107: &\ll  q^{\frac{n}{2}}b^{\frac{1+\sigma+\ve}{2}}\mathcal{S}(\kappa V,a).
1108: \end{align*}
1109: We would now like a version of \cite[Lemma $16$]{41} which
1110: applies to singular forms as well.  We claim that 
1111: \begin{equation}
1112:   \label{eq:JB'} 
1113: \mathcal{S}(\kappa V,a)\ll 
1114: c^\ve d^{\frac{1+\sigma}{2}}V^{n}\Big(1+\frac{c^2d}{V^3}\Big)^{\frac{n}{2}}.
1115: \end{equation}
1116: This relies completely on first establishing suitable analogues of
1117: \cite[Lemmas 13 and 14]{41}.  
1118: A little thought reveals that in the present setting we have
1119: $$
1120: \sum_{|\mathbf{r}|\leq R}N_m(\mathbf{r})^{\frac{1}{2}}\ll
1121: m^{\frac{n}{2}}\Big(1+\frac{R^3}{m}\Big)^{\frac{n}{2}}R^\ve,
1122: $$
1123: for any $m\in  \NN$ and $R\geq 1$.
1124: Here we have used the fact that $C$ is good to bound the number of
1125: $|\mathbf{r}|\leq R$ such that $\rank M(\mathbf{r})=t$, rather than
1126: using \cite[Lemma 2]{41},  as there. Furthermore, we have
1127: $$
1128: \sum_{\mathbf{a}\bmod d}N_d(\mathbf{a}) \ll d^{n+1+\sigma+\ve}.
1129: $$
1130: When $c<V$ it follows from the latter
1131: bound and an application of Cauchy's inequality
1132: that $\mathcal{S}(\kappa V,a)\ll d^{(1+\sigma+\ve)/2}V^n$, which is
1133: acceptable for \eqref{eq:JB'}. In the
1134: alternative case, when $c\geq V$, the necessary modifications to the 
1135: proof of \cite[Lemma $16$]{41} are straightforward and we omit full details here.
1136: 
1137: We may now insert \eqref{eq:JB'} into the preceeding estimate for
1138: $\mathcal{T}(V)$ to conclude that 
1139: $$
1140: \mathcal{T}(V)
1141: \ll  q^{\frac{n+1+\sigma+\ve}{2}}
1142: V^{n}\Big(1+\frac{q}{V^3}\Big)^{\frac{n}{2}}.
1143: $$
1144: If $q^{1/3}\leq V$ then this is clearly satisfacory for \eqref{eq:JB}. 
1145: Alternatively, if $V<q^{1/3}$ then we can only enlarge our bound for
1146: $\mathcal{T}(V)$ if we replace $V$ by $q^{1/3}$. But then 
1147: $\mathcal{T}(V)$ is easily seen to be bounded by \eqref{eq:JB} in this
1148: case too. This therefore completes the proof of \eqref{eq:JB}.
1149: 
1150: 
1151: 
1152: 
1153: 
1154: 
1155: 
1156: 
1157: 
1158: Our final task is to establish the second part of the lemma. Since $C$ is good, 
1159: we may combine  Lemma \ref{lem:paris} with 
1160: \eqref{eq:bound_99} to deduce that 
1161: $$
1162: S^*(\al)\ll q^{-\frac{n}{6}} P^n(\log P)^{7n} 
1163: \min \{1, (|\theta|P^3)^{-1}\}^{\frac{n-1}{2}}.
1164: $$
1165: Let us write $T=q^{-B}P^{-3+C}$ for convenience. It therefore follows that
1166: \begin{align*}
1167: \int_{\A} |S^*(\al)|^k\d\al 
1168: &\ll P^{kn+\frac{\ve}{2}}\sum_{q\leq P^A}
1169: q^{1-\frac{kn}{6}} \int_{-T}^{T} 
1170: \min \{1, (|\theta|P^3)\}^{-\frac{k(n-1)}{2}}\d\theta\\
1171: &\ll P^{kn-3+\frac{\ve}{2}}\sum_{q\leq P^A}
1172: q^{1-\frac{kn}{6}} \\
1173: &\ll P^{kn-3+\ve},
1174: \end{align*}
1175: since $kn\geq 12$ and $k\leq 9$.
1176: \end{proof}
1177: 
1178: 
1179: We remark that when $\sigma=0$ it seems likely that an even sharper
1180: error term is available in Lemma \ref{lem:F} through a more careful
1181: analysis of the complete exponential sums $S_{a,q}(\v)$, when $q$ is prime.
1182: It follows from \cite[Lemma 28]{hooley2} that the form 
1183: $C$ is automatically good when the corresponding hypersurface has at most isolated
1184: singularities and these are suitably mild.
1185: 
1186: In the setting of $1$-dimensional exponential sums, we have even
1187: better control over $S(\al)$ on the arcs $\A=\A(A,B,C)$.  Let $C(x)=cx^3$ for some non-zero coefficient $c\in \ZZ$.  Then for
1188: any $a,q\in \ZZ$ such that $0\leq a< q\leq P^A$ and $\gcd(a,q)=1$, and
1189: any $\al=a/q+\theta\in \A_{q,a}$, the standard major arc analysis would
1190: provide an estimate of the shape
1191: $$
1192: S(\al)= S^*(\al)+O(P^{A}+P^{A+C-AB}),
1193: $$
1194: where $S^*(\al)$ is given by \eqref{eq:S*}. 
1195: Our final result in this section improves on this substantially,
1196: and is readily derived from the book of Vaughan \cite[\S 4]{vaughan}.
1197: 
1198: 
1199: \begin{lemma}\label{lem:E}
1200: Let $\ve>0$, let $n=1$ and let $w\in\WW_1^{(2)}$. 
1201: Let $A,B,C\geq 0$ with $A,B\leq 1$. Then for any $\al\in
1202: \A_{q,a}$ we have 
1203: $$
1204: S(\al)=S^*(\al)+O(P^{\frac{A}{2}+\ve}+P^{\frac{A+C-AB}{2}+\ve}).
1205: $$
1206: Furthermore, if $k\geq 4$, then we have 
1207: $$
1208: \int_\A |S^*(\al)|^k \d\al\ll P^{k-3+\ve}.
1209: $$
1210: \end{lemma}
1211: 
1212: 
1213: 
1214: 
1215: 
1216: 
1217: \section{Density of rational points on cubic hypersurfaces}\label{sec:upper}
1218: 
1219: 
1220: Let $X \subset \PP^{n-1}$ be a cubic hypersurface, not equal to a
1221: cone, that is defined by an absolutely irreducible cubic form $F\in
1222: \ZZ[x_1,\ldots,x_n]$. For $P\geq 1$, let
1223: $$
1224: N_{n,F}(P):=\#\{\x\in\ZZ^n: |\x|\leq P, ~F(\x)=0\},
1225: $$
1226: According to the conjecture of Manin \cite{f-m-t} one expects 
1227: $N_{n,F}(P)\sim cP^{n-3}$ for some constant $c\geq 0$ as soon as $F$ is
1228: non-singular and $n\geq 5$. 
1229: When $F$ is not necessarily non-singular, or the number of variables
1230: is small, there is the dimension growth conjecture due to Heath-Brown. This
1231: predicts that
1232: \begin{equation}
1233:   \label{eq:upper_2}
1234: N_{n,F}(P)\ll P^{n-2+\ve},
1235: \end{equation}
1236: and has received a great deal of attention in recent years.
1237: Let $\sigma$ denote the projective dimension of the singular
1238: locus of $X$. The dimension growth  conjecture has been established by the author
1239: \cite{cubhyp-circle} when $n\geq 6+\sigma$.
1240: The following result, which may of independent interest,  
1241: shows that one can do better than 
1242: \eqref{eq:upper_2} if larger values of $n$ are permitted. 
1243: 
1244: 
1245: \begin{lemma}\label{lem:8}
1246: We have $N_{n,F}(P)\ll P^{n-5/2+\ve}$ when $n\geq 9+\sigma$.
1247: \end{lemma}
1248: 
1249: \begin{proof}
1250: Our proof of the lemma is based on the approach developed in \cite{cubhyp-circle}.
1251: Arguing with hyperplane sections, as in \cite[\S 2]{cubhyp-circle}, we
1252: see that it will suffice to show that there is an absolute constant
1253: $\theta>0$ such that 
1254: \begin{equation}
1255:   \label{eq:gen1}
1256: N_{w}(g;P)
1257: := \sum_{\substack{\x\in \ZZ^n\\ g(\x)=0}} w(P^{-1}\x)
1258: \ll H^{\theta}P^{n-\frac{5}{2}+\ve},
1259: \end{equation}
1260: for any weight function $w:\RR^n\rightarrow \RR_{\geq 0}$ belonging to
1261: the class of weight functions described at the start of \cite[\S 2]{cubhyp-circle}, any 
1262: $H \geq \|g\|_P$, and any 
1263: cubic polynomial $g\in  \ZZ[x_1,\ldots,x_n]$ such that $n\geq 8$ and 
1264: the cubic part $g_0$ is non-singular.
1265: Here we recall that $\|g\|_P:= \|P^{-3}g(P\x)\|$, where $\|h\|$
1266: denotes the height of a polynomial $h$. 
1267: 
1268: The bulk of \cite{cubhyp-circle} goes through
1269: verbatim, and we are left with reevaluating the 
1270: estimation of $\Sigma_2=\Sigma_2(R,\mathbf{R};t)$ and $\Sigma_1=\Sigma_1(R,\mathbf{R};t)$ in 
1271: \cite[\S 5.1]{cubhyp-circle} and \cite[\S 5.2]{cubhyp-circle}, respectively.
1272: Beginning with the former, 
1273: we note from \cite[Eq. (5.5)]{cubhyp-circle} 
1274: that this breaks into an estimation of $\Sigma_{2,a}$ and $\Sigma_{2,b}$.
1275: The first of these is estimated as $O(H^{\theta}
1276: P^{3n/4-3/4+\ve}+H^{\theta} P^{n-3+\ve})$. Both of the exponents of
1277: $P$ are clearly at most $n-5/2+\ve$ when $n\geq 8$, as required for
1278: \eqref{eq:gen1}.
1279: Turning to $\Sigma_{2,b}$, one easily traces through the argument,
1280: finding that
1281: $$
1282: \Sigma_{2,b}\ll H^\theta P^\ve \big(
1283: P^{\frac{3n}{4}-\frac{3}{4}}+ P^{n-2}E_n + P^{\frac{13n}{16}-1}+P^{\frac{7n}{8}-\frac{5}{3}} 
1284: \big),
1285: $$
1286: where 
1287: $$
1288: E_n=P^{-1-\frac{7n}{40}}R^{2-\frac{3n}{20}}R_2^{\frac{3n}{10}-\frac{3}{2}}\ll 
1289: P^{-1-\frac{7n}{40}}R^{\frac{5}{4}}\ll  P^{\frac{7}{8}-\frac{7n}{40}}.
1290: $$
1291: Here the first term (resp. second term, sum of the final two terms) corresponds to the case $V\geq R_2$
1292: (resp. $(R_2^2R_3)^{1/3}\leq V<R_2$, $V<(R_2^2R_3)^{1/3}$). 
1293: A modest pause for thought reveals that all of these exponents are
1294: satisfactory when $n\geq 8$.
1295: 
1296: We now turn to the estimation of $\Sigma_1$ in 
1297: \cite[\S 5.2]{cubhyp-circle}, which is again written as a sum 
1298: $\Sigma_{1,a}+\Sigma_{1,b}$.  Beginning with $\Sigma_{1,a}$, we easily
1299: observe that
1300: $$
1301: \Sigma_{1,a}\ll H^\theta P^\ve \big(
1302: P^{\frac{3n}{4}-\frac{3}{4}}+ P^{n-2}E_n + P^{n-3}\big),
1303: $$
1304: this time with 
1305: $$
1306: E_n=P^{2-\frac{n}{4}}t^{1-\frac{n}{12}}R^{\frac{11}{6}-\frac{n}{4}}(R_2^2R_3)^{\frac{n}{9}-\frac{1}{2}}
1307: \ll P^{-\frac{1}{2}}R^{-\frac{1}{9}}+P^{-1}R^{\frac{2}{9}}\ll P^{-\frac{1}{2}}.
1308: $$
1309: This therefore shows that $\Sigma_{1,a}\ll H^\theta P^{n-5/2+\ve}$, as
1310: required for \eqref{eq:gen1}.  Turning to 
1311: $\Sigma_{1,b}$,  we will need to modify  the argument slightly.
1312: On noting that $R_2^{3/2}R_3^{1/2}\gg (R_2^2R_3)^{2/3}$,
1313: we easily deduce that 
1314: $$
1315: \Sigma_{1,b}\ll H^\theta P^\ve \big(
1316: P^{\frac{3n}{4}-\frac{3}{4}}+ P^{n-2+\frac{7}{8}-\frac{7n}{40}} + T\big),
1317: $$
1318: where we have set
1319: $$
1320: T:=P^nt R^{2-\frac{n}{2}} (R_2^2 R_3)^{-\frac{2}{3}}
1321: \min\Big\{R_2^n, 
1322: \Big(\frac{R_2^2R_3}{V}\Big)^{\frac{n}{2}}, R^{\frac{3n}{8}}\min\{1, (tP^3)^{-\frac{n}{8}}
1323: \Big\}.
1324: $$
1325: The first and second terms here are satisfactory for $n\geq 8$.
1326: Moreover the third term is clearly satisfactory for $n\geq 16$, on
1327: taking $\min\{A,B,C\}=C$. To handle the contribution from the final term when $8\leq n<16$, it will be convenient
1328: to recall that $V$ has order of magnitude $Rt^{1/2}P^{1/2}$ when
1329: $t\geq P^{-3}$ and $R/P$ when $t<P^{-3}$. 
1330: 
1331: Suppose first that $R\geq
1332: P$. When $t\geq P^{-3}$ we deduce that 
1333: \begin{align*}
1334: T\ll P^{n-3} \min\big\{ 
1335: R^{2-n}(R_2^2R_3)^{\frac{n}{2}-\frac{2}{3}}P^{\frac{n}{2}},
1336: R^{2-\frac{n}{8}} (R_2^2R_3)^{-\frac{2}{3}}\big\}
1337: \ll P^{n-\frac{7}{3}}R^{\frac{5}{6}-\frac{n}{8}}
1338: &\ll P^{\frac{7n}{8}-\frac{3}{2}},
1339: \end{align*}
1340: on taking 
1341: \begin{equation}
1342:   \label{eq:ab}
1343:   \min\{A,B\}\leq A^{\frac{4}{3n}}B^{1-\frac{4}{3n}}.
1344: \end{equation}
1345: This is clearly satisfactory for $n\geq 8$. When $t<P^{-3}$ we easily deduce that the
1346: same bound holds on taking $V$ to be of size $R/P$ in the definition
1347: of $T$.
1348: 
1349: Suppose now that $R<P$ and $t\geq P^{-3}$. 
1350: If $t>(R^2P)^{-1}$, then it is not hard to see that
1351: \begin{align*}
1352: T\ll P^{n-1} R^{-\frac{n}{2}} (R_2^2 R_3)^{-\frac{2}{3}}
1353: \min\big\{(R_2^2R_3)^\frac{n}{2}, R^{\frac{5n}{8}}P^{-\frac{n}{4}}\big\}
1354: \ll P^{\frac{3n}{4}-\frac{2}{3}}R^{\frac{n}{8}-\frac{5}{6}}
1355: &\ll P^{\frac{7n}{8}-\frac{3}{2}},
1356: \end{align*}
1357: using \eqref{eq:ab}.  This is satisfactory for $n\geq
1358: 8$. Alternatively, if $P^{-3}\leq 
1359: t\leq (R^2P)^{-1}$, then one finds that
1360: \begin{align*}
1361: T&\ll P^{n} R^{2-\frac{n}{2}} (R_2^2 R_3)^{-\frac{2}{3}}
1362: \min\big\{t(R_2^2R_3)^\frac{n}{2},
1363: R^{\frac{3n}{8}}t^{1-\frac{n}{8}}P^{-\frac{3n}{8}}\big\}.
1364: \end{align*}
1365: Using \eqref{eq:ab} it easily follows that
1366: \begin{align*}
1367: T&\ll P^{\frac{5n}{8}+\frac{1}{2}} t^{\frac{7}{6}-\frac{n}{8}}R^{\frac{3}{2}-\frac{n}{8}}.
1368: \end{align*}
1369: Since $t\leq (R^2P)^{-1}$ and $R<P$ this is clearly satisfactory when
1370: $n=8$.  If instead $n\geq 9$ then we deduce that
1371: \begin{align*}
1372: T&\ll P^{n-3} R^{5-\frac{n}{2}}(R_2^2 R_3)^{\frac{n}{2}-\frac{14}{3}},
1373: \end{align*}
1374: on taking
1375: $\min\{A,B\}\leq A^{1-\frac{8}{n}}B^{\frac{8}{n}}$ rather than \eqref{eq:ab}.
1376: Since $R<P$ we easily conclude that 
1377: $T\ll P^{n-5/2}$ in this case too.
1378: Finally, when $R<P$ and $t<P^{-3}$, we see that
1379: \begin{align*}
1380: T\ll P^{n-3}
1381: \min\big\{ R^{2-\frac{n}{2}}(R_2^2R_3)^{\frac{n}{2}-\frac{2}{3}}, R^{2-\frac{n}{8}}(R_2^2R_3)^{-\frac{2}{3}} 
1382: \big\} \ll P^{n-3} R^{\frac{3}{2}-\frac{n}{8}}
1383: &\ll P^{n-\frac{5}{2}},
1384: \end{align*}
1385: using \eqref{eq:ab}. This therefore concludes the proof of the lemma.
1386: \end{proof}
1387: 
1388: 
1389: It is clear from the proof of Lemma \ref{lem:8} that one actually
1390: achieves an estimate of the shape 
1391: $$
1392: N_{n,F}(P)\leq c_{\ve,n}\|F\|^\theta P^{n-5/2+\ve},
1393: $$ 
1394: for a constant
1395: $\theta>0$, when $n\geq 9+\sigma$.
1396: It seems likely that one can push the analysis further, obtaining
1397: $N_{n,F}(P)\ll P^{n-3+\ve}$ for $n\geq 11+\sigma$, as predicted by Manin.
1398: 
1399: A key step in our argument involves generating good estimates for the
1400: moments 
1401: \begin{equation}
1402:   \label{eq:moment}
1403:   M_{n}(P):=\int_{0}^1|S(\al)|^{2}\d\al,
1404: \end{equation}
1405: where $S(\al)$ is the cubic exponential sum \eqref{eq:Tcubic}, for
1406: an appropriate weight $w\in\WW_n$. 
1407: By the orthogonality of the exponential
1408: function we have 
1409: $$
1410: M_n(P) =\sum_{\substack{\x,\y\in\ZZ^n\\C(\x)=C(\y)}} w(P^{-1}\x)w(P^{-1}\y).
1411: $$
1412: It is clear that there exists a constant $c>0$ depending on $w$ such
1413: that the overall contribution to $M_n(P)$
1414: from $\x,\y$ such that $\max\{|\x|, |\y|\} >c P$ is
1415: $O(1)$,  if $P$ is taken to be sufficiently large. Hence
1416: it follows that
1417: $$
1418: M_n(P)\ll N_{2n,C-C}(cP).
1419: $$
1420: When $C$ is a non-singular form in $n$ variables it is obvious that $C-C$
1421: is a non-singular form in $2n$ variables, defining a hypersurface of
1422: dimension $2n-2$.  
1423: When $C$ has a finite non-empty singular locus it is not hard to see
1424: that $C-C$ has singular locus of dimension $1$.
1425: The following result now flows very easily from
1426: \eqref{eq:upper_2} and Lemma \ref{lem:8}.
1427: 
1428: 
1429: \begin{lemma}\label{lem:C}
1430: Let $\ve>0$ and let $n\geq 3$.
1431: Assume that $C\in \ZZ[x_1,\ldots,x_n]$ is a cubic form defining a
1432: projective hypersurface whose singular locus has dimension $\sigma\in\{-1,0\}$.
1433: Then we have  
1434: $$
1435: M_n(P) \ll 
1436: \begin{cases}
1437: P^{4+\ve},
1438: &\mbox{if $n=3$ and $\sigma=-1$,}\\
1439: P^{2n-\frac{5}{2}+\ve}, &\mbox{if $n\geq 5+\sigma$.}
1440: \end{cases}
1441: $$
1442: \end{lemma}
1443: 
1444: 
1445: 
1446: 
1447: 
1448: 
1449: 
1450: 
1451: 
1452: 
1453: 
1454: 
1455: 
1456: 
1457: 
1458: 
1459: 
1460: 
1461: 
1462: 
1463: \section{Cubics splitting off a form}
1464: 
1465: 
1466: In this section we establish Theorem \ref{main}. 
1467: Let $n_1,n_2\geq 1$ such that 
1468: $$
1469: n_1+n_2=n\geq 13.
1470: $$ 
1471: It will be convenient to write $\x=(x_1,\ldots,x_{n_1})$ and $\y=(y_1,\ldots,y_{n_2})$.
1472: We henceforth fix our attention on cubic forms of the shape 
1473: $$
1474: C(\x,\y)=C_1(\x)+C_2(\y),
1475: $$
1476: with $C_1\in \ZZ[\x]$ and $C_2\in \ZZ[\y]$.  
1477: In what follows we may always suppose that $C=C_1+C_2$ is non-degenerate, by which we mean that it
1478: is not equivalent over $\ZZ$ to a cubic form in fewer variables, since
1479: such forms have obvious non-zero integral solutions.
1480: Recall the definition of ``good'' cubic forms from \S \ref{s:geom}.
1481: It follows from \cite[\S 14]{dav-book} that either $C_1$ is good or else the cubic hypersurface $C_1=0$ has a
1482: rational point. The same is true for the cubic forms $C_2$ and
1483: $C_1+C_2$. Since the existence of a rational point on any of these
1484: hypersurfaces is enough to ensure that $X(\QQ)\neq \emptyset$ in the
1485: statement of Theorem \ref{main}, so we may proceed under the assumption that
1486: $C_1, C_2$ and $C_1+C_2$ are all good.
1487: 
1488: Let $w=(w_1,w_2)\in \WW_n$, as introduced in \S \ref{sec:exp_sums}.
1489: When $C_1$ satisfies the hypotheses in Lemma \ref{lem:F} we will
1490: assume that $w_1\in\WW_{n_1}^{(1)}$ is the weight function
1491: constructed there.  
1492: Our argument revolves around establishing an asymptotic formula for
1493: the sum
1494: $$
1495: N(P):=\sum_{\substack{(\x,\y)\in \ZZ^n\\ C_1(\x)+C_2(\y)=0}} 
1496: w_1(P^{-1}\x)w_2(P^{-1}\y),  
1497: $$
1498: as $P\rightarrow \infty$.  
1499: As is usual in applications of the Hardy--Littlewood circle method,
1500: the starting point is the simple identity
1501: $$
1502: N(P)=\int_0^1 S_1(\al)S_2(\al) \d\al,  
1503: $$
1504: where 
1505: $$
1506: S_i(\al):=\sum_{\x\in\ZZ^{n_i}}w_i(P^{-1}\x) e(\al C_i(\x))  
1507: $$
1508: for $i=1,2$. It will be convenient to define 
1509: $$
1510: S(\al):=S_1(\al)S_2(\al)=\sum_{(\x,\y)\in\ZZ^{n}}w(P^{-1}(\x,\y))
1511: e\big(\al (C_1(\x)+C_2(\y))\big),
1512: $$
1513: where $w=(w_1,w_2)$.  Then $S(\al)=S_w(\al;C_1+C_2,P)$
1514: is a cubic exponential sum of the sort introduced in \eqref{eq:Tcubic}.
1515: 
1516: 
1517: In the usual way one divides the interval $[0,1]$ into a
1518: set of major arcs and minor arcs.  For major arcs we  will take 
1519: the union of intervals 
1520: $$
1521: \M:=\bigcup_{q\leq P^{\D}}\bigcup_{\substack{a=0\\ \gcd(a,q)=1}}^{q-1} \Big[\frac{a}{q}-
1522: P^{-3+\D}, \frac{a}{q}+P^{-3+\D}
1523: \Big],
1524: $$
1525: which is equal to $\A(\D,0,\D)$ in the notation of \eqref{eq:Aqa}.
1526:  The corresponding set of minor arcs is defined modulo $1$ as
1527: $\m=[0,1]\setminus \M$.
1528: Here $\D>0$ is an arbitrary small parameter. It turns out the choice 
1529: $$
1530: \D:=\frac{1}{10}
1531: $$
1532: is acceptable.  
1533: We may deduce from Lemma~15.4 and \S\S 16--18 in \cite{dav-book} 
1534: (see also \cite[Lemma~2.1]{14}) that
1535: $$
1536: \int_\M S(\al)\d\al = \ss \ii P^{n-3}+o(P^{n-3}),
1537: $$
1538: where 
1539: \begin{align*}
1540: \ss&:=\sum_{q=1}^\infty 
1541: \sum_{\substack{a \bmod{q}\\ \gcd(a,q)=1}} 
1542: q^{-n}S_{a,q}^{(1)}S_{a,q}^{(2)},\\
1543: \ii&:=\int_{-\infty}^{\infty}
1544: \int_{\RR^{n_1}}\int_{\RR^{n_2}}w(\x,\y)e\big(\theta(C_1(\x)+C_2(\x))\big)\d\x\d\y
1545: \d \theta
1546: \end{align*}
1547: are both absolutely convergent.
1548: Here, the absolute convergence of $\ss$ follows from Lemma
1549: \ref{lem:paris}, and  we have written
1550: $$
1551: S_{a,q}^{(i)}:=\sum_{\mathbf{u}\in (\ZZ/q\ZZ)^{n_i}} e_q\big(aC_i(\mathbf{u})\big),
1552: $$
1553: for $i=1,2$.
1554: Since $\ss$ is absolutely convergent and $C=C_1+C_2$ is
1555: non-degenerate, it follows from standard arguments (see \cite[Lemma 7.3]{dav-32}, for example)
1556: that $\ss>0$. The treatment of the singular integral is routine and we omit giving the details
1557: here, all of which can be supplied by consulting 
1558: \cite[\S 16]{dav-book} and \cite[\S 4]{hb-10}. Assuming that 
1559: neither $C_1$ nor $C_2$ has a linear factor defined over
1560: $\QQ$ it is possible to choose 
1561: $(\z_1,\z_2)\in \RR^{n}$ in the definition of
1562: $w=(w_1,w_2)$, so that 
1563: each $\z_i$ is a non-singular real solution to $C_i=0$. On 
1564: selecting a sufficiently small value of $\rho>0$ in the definition
1565: of $w_2$ we can then ensure $\ii> 0$. 
1566: The case in which $C_1$ or $C_2$ does factorise over $\QQ$ 
1567: clearly enables us to deduce the statement of Theorem \ref{main}
1568: very easily.
1569: 
1570: In order to conclude the proof of
1571: Theorem~\ref{main} it remains to show that the overall contribution
1572: from the minor arcs 
1573: \begin{equation}
1574:   \label{eq:minor}
1575: E:=\int_{\m} S_1(\al)S_2(\al) \d \al,
1576: \end{equation}
1577: is satisfactory.  This is
1578: where the bulk of our work lies and we will find it necessary to
1579: undertake a lengthy case by case analysis to handle the 
1580: different values of $n_1$ and
1581: $n_2$. In doing so it will suffice to handle the case $n_1+n_2=13$,
1582: the case $n_1+n_2>13$ being taken care of by \cite{14}.
1583: Without loss of generality we assume henceforth that $1\leq n_1\leq 6$.
1584: 
1585: 
1586: 
1587: Let $Q\geq 1$ and let $\al \in \m$. By Dirichlet's approximation theorem
1588: we may find coprime integers $1 \leq a \leq q$ such that $q \leq Q$
1589: and $|q\al-a|\leq 1/Q$.   The value of $Q$ should satisfy $1\leq Q
1590: \leq P^{3/2}$ and is chosen to optimise the final stages of the argument.
1591: The obvious approach involves applying estimates for each individual
1592: exponential sum $S_1(\al)$ and $S_2(\al)$ for $\al\in \m$, before then
1593: deriving an estimate for the integral over the full set of minor
1594: arcs. While we have rather good control over these sums when $n_1$ and
1595: $n_2$ are both large,
1596: the case in which one of $n_1$ or $n_2$ is small presents more of an
1597: obstacle.  Instead we apply H\"older's
1598: inequality to deduce that
1599: \begin{equation}
1600:   \label{eq:split}
1601:   |E| \leq \left(\int_{\m} |S_1(\al)|^u \d \al \right)^{\frac{1}{u}}\left(\int_{\m} |S_2(\al)|^v \d \al \right)^{\frac{1}{v}},
1602: \end{equation}
1603: for any $u,v>0$ such that $1/u+1/v=1$. This will allow us to separate
1604: out the behaviour of the exponential 
1605: sums $S_1(\al)$ and $S_2(\al)$ on the minor arcs.
1606: 
1607: 
1608: Before embarking on the case by case analysis alluded to above, it
1609: will save needless repetition if we give some reasonably general
1610: estimates here that can be applied in various contexts.
1611: Our principal means for dealing with small values of $n_1$ relies on
1612: taking the inequality 
1613: \begin{equation}
1614:   \label{eq:Ivu}
1615: \int_{\m} |S_1(\al)|^u \d \al \leq \int_{0}^1 |S_1(\al)|^u \d \al
1616: \end{equation}
1617: in \eqref{eq:split}.  This will in turn be estimated as $O(P^{k+\ve})$
1618: for an appropriate $k>0$, whence a typical scenario 
1619: entails studying
1620: \begin{equation}
1621:   \label{eq:Iuv}
1622:   I_{u,v}(k;\mathfrak{n}):=P^{\frac{k}{u}+\ve}\left(\int_{\mathfrak{n}}|S_2(\al)|^v \d\al\right)^{\frac{1}{v}},
1623: \end{equation}
1624: for $u,v>0$ such that $1/u+1/v=1$ and certain subsets
1625: $\mathfrak{n}\subseteq\m$. We will always assume that $6/5< v\leq 2$.
1626: 
1627: 
1628: Let $\al \in \mathfrak{n}$ and let $Q\geq 1$.  There exist
1629: coprime integers $0 \leq a <q \leq Q$ such that 
1630:  $|q\al-a|\leq 1/Q$.  An argument based on dyadic summation reveals that 
1631: \begin{equation}
1632:   \label{eq:foot}
1633:   I_{u,v}(k;\mathfrak{n})\ll P^{\frac{k}{u}+\ve} (\log P)^2 \max_{R,\phi,\pm} \Sigma_v(R,\phi,\pm)^{\frac{1}{v}}, 
1634: \end{equation}
1635: where $\Sigma_v(R,\phi,\pm)$ is given by \eqref{eq:sig}, and 
1636: the maximum is over the possible sign changes and 
1637: $R,\phi$ such that 
1638: \begin{equation}
1639:   \label{eq:range}
1640: 0<R\leq Q, \quad 0<\phi\leq (RQ)^{-1}.
1641: \end{equation}
1642: Furthermore, $R,\phi$ should satisfy whatever conditions are
1643: appropriate to ensure we are dealing with points on $\mathfrak{n}$. In
1644: particular, since $\mathfrak{n}\subseteq \m$  the inequalities
1645: $R\leq P^\D$ and $\phi \leq P^{-3+\D}$ cannot both
1646: hold simultaneously. 
1647: 
1648: Let $u,v,k$ be given. Define
1649: \begin{equation}
1650:   \label{eq:rp1}
1651:   \rho_n:=\frac{n(vn-8-v)}{vn^2-(3v+4)n+2v}, \quad \pi_n:=
1652: \frac{-2v(n^2-(18-2\frac{k}{u})n-2)}{vn^2-(3v+4)n+2v},
1653: \end{equation}
1654: and 
1655: \begin{equation}
1656:   \label{eq:rp2}
1657:   \rho_n':=\frac{2v}{2-v}\left(\frac{n^2}{2(3n-2)}-\frac{2}{v}\right), \quad \pi_n':=
1658: \frac{2v}{2-v}\left(n-\frac{23}{2}+\frac{k}{u}\right).
1659: \end{equation}
1660: Let
1661: \begin{equation}
1662:   \label{eq:delta}
1663: \delta:=\frac{1}{10^4}.
1664: \end{equation}
1665: Recall that our task is to show that
1666: $E=o(P^{10})$ when $n=n_1+n_2=13$.
1667: The following result provides us with easily checked
1668: conditions on $u,v,k$ and $n_2$ under which 
1669: $I_{u,v}(k;\mathfrak{n})$ makes a satisfactory contribution.
1670: 
1671: \begin{lemma}\label{lem:v<2}
1672: Let $6/5< v<2$. Assume that $n_2\geq 6$ and 
1673: \begin{equation}
1674:   \label{eq:paris}
1675:   \rho_{n_2}+\rho_{n_2'}\geq 1.
1676: \end{equation}
1677: Define $\m_0$ to be the set 
1678: of $\al \in \m$ for which there exist coprime integers $0 \leq a <q$ such that 
1679: $$
1680: q\leq P^{\frac{\pi_{n_2}'-\pi_{n_2}}{\rho_{n_2}'+\rho_{n_2}}+2\delta}, \quad
1681: q^{\rho_{n_2}'}P^{-\pi_{n_2}'-\delta}\leq 
1682: \Big|\al-\frac{a}{q}\Big|\leq q^{-\rho_{n_2}}P^{-\pi_{n_2}+\delta}.
1683: $$
1684: Then 
1685: $$
1686: I_{u,v}(k;\mathfrak{n}\setminus \m_0)=o(P^{10}),
1687: $$
1688: for any $\mathfrak{n}\subseteq \m$,
1689: provided that 
1690: \begin{equation}\label{eq:in1}
1691: \frac{2}{v}+\frac{21}{2}-n_2 \leq \frac{k}{u}< \frac{103}{10}-\frac{4n_2}{5}
1692: \end{equation}
1693: and 
1694: \begin{equation}
1695:   \label{eq:in2}
1696: \pi_{n_2}'\geq 3.
1697: \end{equation}
1698: \end{lemma}
1699: 
1700: \begin{proof}
1701: We will commence under the assumption that $6/5<v\leq 2$,
1702: saving the restriction $v<2$ until later in the argument.
1703: It is clear that $I_{u,v}(k;\mathfrak{n}\setminus \m_0)\leq I_{u,v}(k;\m\setminus \m_0)$.
1704: Let us consider the consequences of applying Lemma \ref{lem:D} in our
1705: estimate \eqref{eq:foot} for $I_{u,v}(k;\m\setminus \m_0)$. Throughout
1706: the proof of Lemma \ref{lem:v<2} we will denote $\m\setminus \m_0$ by
1707: $\mathfrak{a}$ and we will set $n=n_2$. We may deduce from Lemma
1708: \ref{lem:D} and \eqref{eq:Iuv} that
1709: $$
1710: I_{u,v}(k;\mathfrak{a})
1711: \ll P^{\frac{k}{u}+\ve} \left(P^{\frac{3}{v}}+\max_{R,\phi}
1712: R^{\frac{2}{v}}\phi^{\frac{1}{v}-\frac{1}{2}} \Big( \frac{\psi_H
1713:   P^{2n-1}}{H^{n-1}}F\Big)^{\frac{1}{2}} \right),
1714: $$
1715: where $\psi_H$ and $F$ are as in the statement of the lemma
1716: and $H\in [1,P]\cap \ZZ$ is arbitrary. Furthermore the maximum is over
1717: $R,\phi$ such that
1718: \eqref{eq:range} holds with any choice of $Q\geq 1$ that we care
1719: to choose. We write $Q=P^{\kappa}$, with 
1720: \begin{equation}
1721:   \label{eq:kappa}
1722: \kappa:=\frac{3(2n-21+2\frac{k}{u})}{n-1}+3\ve.
1723: \end{equation}
1724: In particular one easily checks that $0\leq \kappa\leq
1725: 3/2$ if \eqref{eq:in1}  holds and $\ve>0$ is sufficiently small.
1726: It follows from \eqref{eq:in1} that $k/u<11/2$ since $n\geq 6$. Hence
1727: the term involving $P^{3/v}$ contributes $O(P^{8+\ve})$, which is satisfactory. 
1728: 
1729: 
1730: Let  us now turn to the contribution from the term involving
1731: $F$ in our estimate for $I_{u,v}(k;\mathfrak{a})$. Define
1732: \begin{equation}
1733:   \label{eq:sock}
1734:   \phi_0:=(R^{-\frac{2}{v}}P^{-(2n-\frac{23}{2}+\frac{k}{u})})^{\frac{2v}{v(n-1)+2}}.
1735: \end{equation}
1736: Then our investigation will be optimised by taking
1737: $$
1738: H:=
1739: \begin{cases}
1740: \lfloor P^{\ve}\max\{1,R^{\frac{2}{v}}\phi^{\frac{1}{v}}P^{n-\frac{21}{2}+\frac{k}{u}}\}^{\frac{2}{n-1}} \rfloor, &\mbox{if  $\phi>\phi_0$,}\\
1741: \lfloor P^{\ve}\max\{1,R^{\frac{2}{v}}\phi^{\frac{1}{v}-\frac{1}{2}}P^{n-\frac{23}{2}+\frac{k}{u}}\}^{\frac{2}{n}}\rfloor, &\mbox{if  $\phi\leq \phi_0$.}
1742: \end{cases}
1743: $$
1744: If we can show that $F\ll 1$ with this
1745: choice of $H$ then we will have 
1746: $$
1747: \frac{P^{n-\frac{1}{2}+\frac{k}{u}+\ve}
1748:   R^{\frac{2}{v}}\phi^{\frac{1}{v}-\frac{1}{2}}\psi_H^{\frac{1}{2}}}{H^{\frac{n-1}{2}}}
1749: F^{\frac{1}{2}}\ll P^{10-\frac{3\ve}{2}},
1750: $$
1751: since $n \geq 6$. We deduce from \eqref{eq:range} that 
1752: \begin{align*}
1753: (R^{\frac{2}{v}}\phi^{\frac{1}{v}}P^{n-\frac{21}{2}+\frac{k}{u}})^{\frac{2}{n-1}}
1754: &\leq P^{\frac{2(n-\frac{21}{2}+\frac{k}{u})}{n-1}},
1755: \quad (R^{\frac{2}{v}}\phi^{\frac{1}{v}-\frac{1}{2}}P^{n-\frac{23}{2}+\frac{k}{u}})^{\frac{2}{n}}
1756: \leq P^{\frac{2(n-10+\frac{k}{u})}{n}}.
1757: \end{align*}
1758: In either case the final exponent of $P$ is less than 1, by
1759: \eqref{eq:in1}.  
1760: Hence $H$ is an integer in the
1761: interval $[1,P]$ and it remains to show that $F\ll 1$ with this choice of
1762: $H$. Recall the definition of $F$ from Lemma \ref{lem:D}.
1763: 
1764: 
1765: Suppose first that $\phi>\phi_0$, with $\phi_0$ given by
1766: \eqref{eq:sock}. Then $\psi_H\ll \phi$ and it follows that
1767: \begin{align*}
1768: RH^3\psi_H
1769: \ll R\phi P^{3\ve}
1770: (1+R^{\frac{2}{v}}\phi^{\frac{1}{v}}P^{n-\frac{21}{2}+\frac{k}{u}})^{\frac{6}{n-1}}
1771: &\ll 1+ P^{-\kappa} P^{(n-\frac{21}{2}+\frac{k}{u})\frac{6}{n-1}+3\ve},
1772: \end{align*}
1773: by \eqref{eq:range}. It follows from our expression \eqref{eq:kappa}
1774: for $\kappa$ that this is $O(1)$.
1775: Turning to the third term in the definition of 
1776: $F$, we see that
1777: \begin{align*}
1778: \frac{H^{n}}{R^{\frac{n}{2}}(P^2\psi_H)^{\frac{n-2}{2}}}\ll 
1779: \frac{P^{n\ve}}{R^{\frac{n}{2}}\phi^{\frac{n-2}{2}}P^{n-2}}(1+R^{\frac{2}{v}}\phi^{\frac{1}{v}}
1780: P^{n-\frac{21}{2}+\frac{k}{u}})^{\frac{2n}{n-1}}.
1781: \end{align*}
1782: The exponent of $\phi$ in the second term is $2n/(v(n-1))-(n-2)/2$, which is negative
1783: since $v> 6/5$ and $n\geq 6$. Hence this
1784: quantity is $O(1)$ provided that $\phi\geq \phi_1$, with 
1785: \begin{equation}
1786:   \label{eq:knee}
1787:   \phi_1:=R^{-\rho_n}P^{-\pi_n+\delta},
1788: \end{equation}
1789: with $\rho_n, \pi_n$ given by \eqref{eq:rp1} and $\delta$ given by \eqref{eq:delta}.
1790: Here, as is customary, we have assumed that $\ve$ is
1791: sufficiently small.
1792: One also checks that taking $\phi\geq \phi_1$ is enough to ensure
1793: that the first term is $O(1)$, in view of the lower bound for $k/u$ in \eqref{eq:in1}.
1794: 
1795: Suppose now that $\phi\leq \phi_0$.  Then $\psi_H\ll (P^2H)^{-1}$ and it follows that
1796: \begin{align*}
1797: RH^3\psi_H
1798: &\ll \frac{R P^{2\ve}}{P^2}
1799: (1+R^{\frac{2}{v}}\phi^{\frac{1}{v}-\frac{1}{2}}P^{n-\frac{23}{2}+\frac{k}{u}})^{\frac{4}{n}}\\
1800: &\ll 1+  
1801: R^{1+\frac{8}{vn}}(RQ)^{-(\frac{1}{v}-\frac{1}{2})\frac{4}{n}}P^{(n-\frac{23}{2}+\frac{k}{u})\frac{4}{n}-2+2\ve}\\
1802: &\ll 1+  
1803: P^{\kappa(1+\frac{4}{n})}P^{2-\frac{46}{n}+\frac{4k}{un}+2\ve},
1804: \end{align*}
1805: since $Q=P^{\kappa}$. It follows from \eqref{eq:in1} and \eqref{eq:kappa} 
1806: that this is $O(1)$.  Thus the second term makes a satisfactory 
1807: contribution in $F$. Turning to the third term, we find that
1808: \begin{align*}
1809: \frac{H^{n}}{R^{\frac{n}{2}}(P^2\psi_H)^{\frac{n-2}{2}}}\ll
1810:  \frac{P^{\frac{(3n-2)\ve}{2}}}{R^{\frac{n}{2}}}
1811: (1+R^{\frac{2}{v}}\phi^{\frac{1}{v}-\frac{1}{2}}P^{n-\frac{23}{2}+\frac{k}{u}})^{\frac{3n-2}{n}}.
1812: \end{align*}
1813: We now make the assumption $6/5<
1814: v<2$. Hence the overall contribution from the second term
1815: is $O(1)$ provided that $\phi\leq \phi_2$, with 
1816: \begin{equation}
1817:   \label{eq:ankle}
1818:   \phi_2:=R^{\rho_n'}P^{-\pi_n'-\delta},
1819: \end{equation}
1820: with $\rho_n', \pi_n'$ given by \eqref{eq:rp2} and $\delta$ given by \eqref{eq:delta}.
1821: Assuming \eqref{eq:in2} we note that if $\phi\leq \phi_2$ and $R\leq P^{n\ve/(3n-2)}$ then we would 
1822: have a point on the major arcs if $\ve$ is sufficiently small in terms
1823: of $\D$, which we have seen to be impossible. Hence the inequality $\phi\leq
1824: \phi_2$ is also enough to ensure that the first term is $O(1)$.
1825: 
1826: 
1827: When $v=2$ the exponent of $\phi$ is zero in the above and we will
1828: have an overall contribution of $O(1)$ unless 
1829: \begin{equation}
1830:   \label{eq:cas2}
1831: R\leq   P^{\frac{(3n-2)(2n-23+k)}{(n^2-6n+4)}+\delta}.
1832: \end{equation}
1833: We will return to this case shortly.
1834: Recall the definitions \eqref{eq:sock}--\eqref{eq:ankle} of $\phi_0,\phi_1,\phi_2$.
1835: It follows from the inequality $\phi_2<
1836: \phi_1$ that $R^{\rho_n'+\rho_n}< P^{\pi_n'-\pi_n+2\delta}$.
1837: We now employ the assumption \eqref{eq:paris} on the size of $\rho_n'+\rho_n$.
1838: Combining the above we conclude that there is an overall contribution of 
1839: $o(P^{10})$ to $I_{u,v}(k;\mathfrak{a})$ from all of the relevant values of
1840: $R,\phi$, apart from those which satisfy the inequalities
1841: $$
1842: R< P^{{\frac{\pi_n'-\pi_n}{\rho_n'+\rho_n}}+2\delta},\quad \phi_2<\phi <\phi_1.
1843: $$
1844: But then the relevant point is forced to lie in the set $\m_0$
1845: that was defined in the statement of the lemma. This is impossible,
1846: and so completes the proof of Lemma~\ref{lem:v<2}.
1847: \end{proof}
1848: 
1849: 
1850: 
1851: Our next result deals with the corresponding case in which $u=v=2$.
1852: In this setting \eqref{eq:rp1} becomes
1853: \begin{equation}
1854:   \label{eq:rp1'}
1855:   \rho_n=\frac{n(n-5)}{n^2-5n+2}, \quad \pi_n=
1856: \frac{-2(n^2-(18-k)n-2)}{n^2-5n+2}.
1857: \end{equation}
1858: Define
1859: \begin{equation}
1860:   \label{eq:r''}
1861:   \rho_n'':=
1862: \frac{(3n-2)(2n-23+k)}{n^2-6n+4}
1863: \end{equation}
1864: and
1865: \begin{equation}\label{eq:psi}
1866: \psi_n:=\rho_n''\left(1+\frac{n}{8}-\frac{(4+n)\rho_n}{8}\right)
1867: +n+\frac{k}{2}-\frac{(4+n)\pi_n}{8},
1868: \end{equation}
1869: for any $k$ and $n$, and recall the definition \eqref{eq:delta} of
1870: $\delta$. 
1871: Then we have the following result.
1872: 
1873: \begin{lemma}\label{lem:v=2}
1874: Let $u=v=2$. Assume that $n_2\geq 6$.
1875: Define $\m_0$ to be the set 
1876: of $\al \in \m$ for which there exist coprime integers $0 \leq a <q$ such that 
1877: $$
1878: q\leq P^{\rho_{n_2}''+\delta}, \quad
1879: \Big|\al-\frac{a}{q}\Big|\leq 
1880: q^{-\frac{n_2-8}{n_2-4}}P^{-\frac{80-5n_2-4k}{n_2-4}+\delta}.
1881: $$
1882: Then 
1883: $$
1884: I_{2,2}(k;\mathfrak{n}\setminus \m_0)=o(P^{10}),
1885: $$
1886: for any $\mathfrak{n}\subseteq \m$, 
1887: provided that \eqref{eq:in1} holds and 
1888: \begin{equation}\label{eq:in3}
1889: \psi_{n_2}\leq 10-\frac{1}{10}.
1890: \end{equation}
1891: \end{lemma}
1892: 
1893: 
1894: \begin{proof}
1895: We continue to write $\mathfrak{a}=\mathfrak{m}\setminus\m_0$ and 
1896: $n=n_2$ throughout the proof, in order to improve the appearance of
1897: our expressions. Our starting point is the proof of Lemma \ref{lem:v<2}, which on
1898: passing to dyadic intervals via
1899: \eqref{eq:foot}, shows that 
1900: $I_{2,2}(k;\mathfrak{a})=o(P^{10})$ 
1901: unless $\phi<\phi_1$, in the notation of \eqref{eq:knee}, 
1902: and  the inequality \eqref{eq:cas2} holds for $R$.
1903: This much is valid subject to \eqref{eq:in1}.
1904: 
1905: We now consider the effect of applying Lemma \ref{lem:B} in
1906: \eqref{eq:foot} when $R$ and $\phi$ are in the remaining ranges, with
1907: $u=v=2$. This gives
1908: \begin{align*}
1909: I_{2,2}(k;\mathfrak{a})
1910: &\ll P^{n+\frac{k}{2}+2\ve} \max_{R,\phi} 
1911: (R^2\phi)^{\frac{1}{2}} \big(R\phi+ (R\phi P^3)^{-1}\big)^{\frac{n}{8}}\\
1912: &\ll  P^{2\ve}\max_{R,\phi} \big(
1913: R^{1+\frac{n}{8}}\phi^{\frac{4+n}{8}}P^{n+\frac{k}{2}} +
1914: R^{1-\frac{n}{8}}\phi^{\frac{4-n}{8}} P^{\frac{5n}{8}+\frac{k}{2}}\big).
1915: \end{align*}
1916: Taking $\phi<\phi_1$ and recalling the assumed inequality
1917: \eqref{eq:cas2} for $R$ we see that the first term here is 
1918: \begin{align*}
1919: &\ll   
1920: R^{1+\frac{n}{8}}(R^{-\rho_n}P^{-\pi_n+\delta})^{\frac{4+n}{8}}P^{n+\frac{k}{2}+2\ve}\\
1921: &\ll   
1922: R^{1+\frac{n}{8}-\frac{(4+n)\rho_n}{8}}
1923: P^{n+\frac{k}{2}-\frac{(4+n)\pi_n}{8}+2\ve+(\frac{4+n}{8})\delta}\\
1924: &\ll   P^{\psi_n+2\ve+(1+\frac{n}{8}-\frac{(4+n)\rho_n}{8}+\frac{4+n}{8})\delta},
1925: \end{align*}
1926: where $\psi_n$ is given by \eqref{eq:psi}. According to  
1927: \eqref{eq:in3} this contribution is satisfactory. 
1928: Turning to the second term in the above estimate for
1929: $I_{2,2}(k;\mathfrak{a})$, we will have $O(P^{10-\ve})$ as an upper
1930: bound for this quantity provided
1931: that $\phi>\phi_3$, with 
1932: $$
1933: \phi_3:=R^{-\frac{n-8}{n-4}}P^{-\frac{80-5n-4k}{n-4}+\delta},
1934: $$
1935: since $n\geq 6$ by assumption. 
1936: 
1937: Our investigation has therefore allowed us to handle all $\al$ apart from
1938: those for which $R\leq P^{\rho_n''+\delta}$ and $\phi<\phi_3$, where
1939: $\rho_n''$ is given by \eqref{eq:r''}.
1940: Such points are forced to lie on the set of arcs defined in $\m_0$. This therefore
1941: completes the proof of Lemma \ref{lem:v=2}.
1942: \end{proof}
1943: 
1944: 
1945: 
1946: The ideal scenario is when we can apply
1947: Lemma \ref{lem:v=2} with 
1948: $$
1949: k=2n_1-3=23-2n_2,
1950: $$
1951: and we will find this is possible for 
1952: certain ranges of $n_1,n_2$ such that $n_1+n_2=13$. When this comes to
1953: pass it follows from \eqref{eq:rp1'}, \eqref{eq:r''} that
1954: $\pi_{n_2}=2$ and $\rho_{n_2}''=0$, and furthermore,
1955: $\psi_{n_2}=21/2-n_2/4$. One easily checks that the conditions in
1956: \eqref{eq:in1} and \eqref{eq:in3} are satisfied for $n_2\geq 6$. 
1957: Finally we note that $\m\setminus \m_0=\m$ in the statement of
1958: Lemma \ref{lem:v=2} since clearly any element of $\m_0$ is forced
1959: to lie on the major arcs. We may conclude as follows. 
1960: 
1961: \begin{lemma}\label{lem:v=2'}
1962: Assume that $n_2\geq 6$.  Then 
1963: we have 
1964: $$
1965: I_{2,2}(2n_1-3;\mathfrak{m})=o(P^{10}).
1966: $$
1967: \end{lemma}
1968: 
1969: 
1970: 
1971: 
1972: We are now ready to apply this collection of estimates in our case by
1973: case analysis of the minor arc integral $E$ in \eqref{eq:minor}.
1974: 
1975: 
1976: 
1977: 
1978: \subsection{The case $n_1=1$}\label{s:1.12}
1979: 
1980: 
1981: We will assume that $w\in \WW_n^{(2)}$ throughout this section. 
1982: One of the ingredients in our treatment of this case 
1983: is the use of ``pruning''. We will find it
1984: convenient to sort the minor arcs into subsets
1985: $$
1986: \emptyset=\mathfrak{n}_3\subseteq \mathfrak{n}_2\subseteq \mathfrak{n}_1
1987: \subseteq \n_0:= \m.
1988: $$ 
1989: Recall the definition \eqref{eq:delta} of $\delta$. We define $\n_1$ to be the set of $\al\in \m$ for
1990: which there exists $a,q\in\ZZ$ such that $0\leq a<q\leq P^{17/24+2\delta}$ and $\gcd(a,q)=1$,
1991: with 
1992: \begin{equation}
1993:   \label{eq:box1}
1994: q^{\frac{42}{17}}P^{-4-\delta}\leq \Big|\al-\frac{a}{q}\Big|\leq q^{-1}P^{-\frac{37}{24}+2\delta},
1995: \end{equation}
1996: We denote by $\n_2$ the corresponding set of $\al \in \n_1$ with the
1997: property that whenever \eqref{eq:box1} holds with $\gcd(a,q)=1$ and $0\leq a< q\leq P^{17/24+\delta}$, then
1998: $$
1999: q\leq P^{\frac{27}{50}}.
2000: $$
2001: We will write $E_i$ for the overall contribution to $E$ from
2002: integrating over the set $\n_i\setminus \n_{i+1}$, for $i=0,1,2$.
2003: Our task is to show that $E_i=o(P^{10})$ for each $i$.
2004: 
2005: To handle the case $i=0$ we begin as in \eqref{eq:split} and
2006: \eqref{eq:Ivu} with $(u,v)=(4,4/3)$.  It easily follows that
2007: $$
2008: \int_{0}^1 |S_1(\al)|^4 \d \al \ll P^{2+\ve},
2009: $$
2010: on interpreting the integral as a sum over the solutions of
2011: the equation $x_1^3+x_2^3=x_3^3+x_4^3$, with $x_i\ll P$, and
2012: applying standard estimates for the divisor function. Hence we have
2013: $$
2014: E_0 \ll I_{4,\frac{4}{3}}(2;\m\setminus \n_1),
2015: $$
2016: in the notation of \eqref{eq:Iuv}.
2017: When $(u,v)=(4,4/3)$, $k=2$ and $n_2=12$ we have 
2018: $$
2019:  \rho_{12}=\frac{30}{37}, \quad \pi_{12}=\frac{62}{37},\quad
2020: \rho_{12}'=\frac{42}{17},\quad 
2021: \pi_{12}'=4,
2022: $$
2023: in \eqref{eq:rp1} and \eqref{eq:rp2}. In particular \eqref{eq:paris}, \eqref{eq:in1} and
2024: \eqref{eq:in2} are satisfied. 
2025: Now it is easily to see  that $\m\setminus\n_1\subset
2026: \n\setminus\m_0$, where $\m_0$ is as in the statement of Lemma~\ref{lem:v<2}, since for $\al\in \m_0$ we have 
2027: \begin{align*}
2028: q^{\frac{42}{17}}P^{-4-\delta}\leq 
2029: \Big|\al-\frac{a}{q}\Big|\leq 
2030: q^{-\frac{30}{37}}P^{-\frac{62}{37}+\delta}=q^{-1}q^{\frac{7}{37}}P^{-\frac{62}{37}+\delta}
2031: &\leq q^{-1}P^{-\frac{37}{24}+2\delta}.
2032: \end{align*}
2033: It therefore follows from Lemma \ref{lem:v<2} that $E_0=o(P^{10})$, as
2034: required.
2035: 
2036: Turning to the case $i=1$, we begin as above with the observation that 
2037: $$
2038: E_1 \ll I_{4,\frac{4}{3}}(2;\n_1\setminus \n_2),
2039: $$
2040: This time we appeal to Lemma \ref{lem:B}. On observing that 
2041: $$
2042: \Big|\al-\frac{a}{q}\Big|\leq q^{-1}P^{-\frac{37}{24}+2\delta}\leq q^{-1}P^{-\frac{3}{2}},
2043: $$
2044: for any $\al \in \n_1$, we deduce from the second part of this result
2045: that
2046: \begin{align*}
2047: E_1 &\ll P^{\frac{25}{2}+2\ve} \max_{R,\phi} (R^2\phi)^{\frac{3}{4}}R^{-\frac{3}{2}}\min
2048: \{1,(\phi P^3 )^{-\frac{3}{2}}\}
2049: \ll P^{8+2\ve} \max_{R,\phi} \phi^{-\frac{3}{4}},
2050: \end{align*}
2051: where the maximum is over all $R,\phi>0$ such that 
2052: $$
2053: P^{\frac{27}{50}}< R\leq P^{\frac{17}{24}+2\delta}, \quad
2054: R^{\frac{42}{17}}P^{-4-\delta}< \phi < R^{-1}P^{-\frac{37}{24}+2\delta}. 
2055: $$
2056: Taking the lower bounds for $\phi$ and $R$ that emerge from these
2057: inequalities therefore implies that
2058: \begin{align*}
2059: E_1 &\ll P^{11+\frac{3\delta}{4}+2\ve} \max_{R} R^{-\frac{63}{34}} =o(P^{10}),
2060: \end{align*}
2061: on recalling that $\delta=10^{-4}$ from \eqref{eq:delta} and $\ve>0$
2062: is arbitrary.
2063: 
2064: 
2065: The key idea in our treatment of $E_2$ is to take
2066: advantage of the fact that we have rather good control of the
2067: $1$-dimensional exponential sum $S_1(\al)$ on suitable sets of ``major
2068: arcs''. Recall the definition \eqref{eq:Aqa} of $\A=\A(A,B,C)$.
2069: We will take $(A,B,C)=(24/50,1,35/24+2\delta)$, whence we may deduce
2070: from Lemma \ref{lem:E} that
2071: $$
2072: S_1(\al)=S_1^*(\al)+O(P^{\frac{35}{48}+\delta+\ve}),
2073: $$
2074: for any $\al \in \A_{a,q}$, where $S_1^*(\al)$ is given by \eqref{eq:S*}.
2075: It follows that
2076: \begin{equation}\label{eq:improper0}
2077: E_2
2078: \ll
2079: \int_{\n_2} |S_1^*(\al)S_2(\al)|\d\al
2080: +P^{\frac{35}{48}+\delta+\ve} \int_{\n_2} |S_2(\al)|\d\al
2081: =I_1+I_2,
2082: \end{equation}
2083: say.  We will show that $I_1$ and $I_2$ are both $o(P^{10})$.
2084: 
2085: 
2086: Let us begin by analysing the first term in this bound. 
2087: Now it follows from the second part of Lemma \ref{lem:E} and H\"older's inequality that 
2088: \begin{align*}
2089: I_1&\ll I_{4,\frac{4}{3}}(1,\n_2),
2090: \end{align*}
2091: in the notation of \eqref{eq:Iuv}.  A straightforward application of
2092: Lemma \ref{lem:v<2} reveals that $I_{4,4/3}(1,\n_2\setminus
2093: \n_*)=o(P^{10})$, where $\n_*$ is the set of $\al \in \n_1$ for which
2094: there exists $a,q\in\ZZ$ such that $0\leq a<q\leq P^{17/48+2\delta}$ and $\gcd(a,q)=1$,
2095: with 
2096: $$
2097: q^{\frac{42}{17}}P^{-3-\delta}\leq \Big|\al-\frac{a}{q}\Big|\leq q^{-\frac{30}{37}}P^{-\frac{68}{37}+\delta}.
2098: $$
2099: To estimate $I_{4,4/3}(1,\n_*)$ we employ the second part of Lemma
2100: \ref{lem:B} in much the same way that we did in our analysis of $E_1$. This implies that
2101: \begin{align*}
2102: I_{4,\frac{4}{3}}(1,\n_*) &\ll P^{\frac{49}{4}+2\ve} \max_{R,\phi} (R^2\phi)^{\frac{3}{4}}R^{-\frac{3}{2}}\min
2103: \{1,(\phi P^3 )^{-\frac{3}{2}}\},
2104: \end{align*}
2105: where the maximum is over all $R,\phi>0$ such that 
2106: $$
2107: R\leq P^{\frac{17}{48}+2\delta}, \quad
2108: R^{\frac{42}{17}}P^{-3-\delta}< \phi <
2109: R^{-\frac{30}{37}}P^{-\frac{68}{24}+\delta}, 
2110: $$
2111: with the inequalities $R\leq P^\D$ and $\phi \leq P^{-3+\D}$ not both
2112: holding simultaneously. 
2113: Taking the lower bound for $\phi$ we obtain the contribution
2114: $$
2115: \ll P^{\frac{31}{4}+2\ve} \phi^{-\frac{3}{4}} 
2116: \ll P^{10+\frac{3\delta}{4}+2\ve}R^{-\frac{63}{34}}.
2117: $$
2118: This is $o(P^{10})$ if $R\geq P^{\delta}$. If on the other
2119: hand $R<P^{\delta}\leq P^{\D}$ we must automatically have $\phi \geq P^{-3+\D}$,
2120:   whence we still obtain a satisfactory contribution. This completes the treatment of 
2121: $I_{4,4/3}(1,\n_*)$, and so that of $I_1$. 
2122: 
2123: 
2124: 
2125: We now turn to the contribution from $I_2$ in \eqref{eq:improper0}.
2126: Breaking the ranges for $q$ and $|\al-a/q|$ into dyadic intervals as
2127: usual, and applying the second part of Lemma \ref{lem:B}, we have
2128: $$
2129: I_2\ll P^{12+\frac{35}{48}+\delta+2\ve} \max_{R,\phi} R^{\frac{1}{2}}\phi \min
2130: \{1,(\phi P^3 )^{-\frac{3}{2}}\},
2131: $$
2132: where the maximum is over all $R,\phi$ such that 
2133: $$
2134: 0<R\leq P^{\frac{27}{50}}, \quad  0<\phi < R^{-1}P^{-\frac{37}{24}+2\delta},
2135: $$
2136: with the inequalities $R\leq P^\D$ and $\phi \leq P^{-3+\D}$ not both
2137: holding simultaneously.  If $\phi >P^{-3}$ then this is 
2138: $$
2139: \ll 
2140: R^{\frac{1}{2}} P^{9+\frac{35}{48}+\delta+2\ve} \ll
2141: P^{10-\frac{1}{1200}+\delta+2\ve}=o(P^{10}), 
2142: $$
2143: whereas if on the other hand $\phi\leq P^{-3}$, then the same basic conclusion holds.
2144: 
2145: Once taken all together, this therefore completes the treatment of the
2146: minor arcs when $(n_1,n_2)=(1,12)$. 
2147: 
2148: 
2149: 
2150: \subsection{The case $n_1=2$}
2151: 
2152: 
2153: We will continue to assume that $w\in \WW_n^{(2)}$ throughout this section. 
2154: In what follows we may assume that
2155: $C_1$ does not take the shape
2156: $a(b_1x_1+b_2x_2)^3$, for integers $a,b_1,b_2$, since otherwise
2157: the resolution of Theorem \ref{main} is trivial.
2158: 
2159: 
2160: Recall the manipulations in \eqref{eq:split} and
2161: \eqref{eq:Ivu}. Taking $(u,v)=(4,4/3)$ it
2162: follows from Lemma \ref{lem:hua_2} that the latter 
2163: inequality is bounded  by $O(P^{5+\ve})$.
2164: Thus we are led to estimate $I_{4,4/3}(5;\m)$, as given by
2165: \eqref{eq:Iuv}.  We clearly have 
2166: $$
2167: \rho_{11}=\frac{44}{57}, \quad \pi_{11}=\frac{103}{57},\quad
2168: \rho_{11}'=\frac{56}{31}, \quad \pi_{11}'=3,
2169: $$
2170: in \eqref{eq:rp1} and \eqref{eq:rp2}.
2171: One easily checks that the conditions \eqref{eq:paris}, \eqref{eq:in1} and
2172: \eqref{eq:in2} in the statement of
2173: Lemma  \ref{lem:v<2} are satisfied with our choice of $u,v,k$ and
2174: $n_2$. Recall the definition \eqref{eq:delta} of $\delta$. 
2175: Define $\m_0$ to be the set 
2176: of $\al \in \m$ for which there exist coprime integers $0 \leq a <q$ such that 
2177: $$
2178: q\leq P^{\frac{31}{67}+2\delta},\quad 
2179: q^{\frac{56}{31}}P^{-3-\delta}\leq \Big|\al-\frac{a}{q}\Big|\leq q^{-\frac{44}{57}}P^{-\frac{103}{57}+\delta}.
2180: $$
2181: Then we may conclude from Lemma \ref{lem:v<2} that 
2182: $
2183: I_{4,\frac{4}{3}}(5;\m\setminus \m_0)=o(P^{10}).
2184: $
2185: 
2186: It remains to deal with the contribution from $\al\in \m_0$.
2187: We will use Lemma~\ref{lem:B} to handle this remaining range. 
2188: Now it follows that
2189: $$
2190: \Big|\al-\frac{a}{q}\Big|\leq
2191: q^{-1}q^{\frac{13}{57}}P^{-\frac{103}{57}+\delta}<q^{-1}P^{-\frac{3}{2}}.
2192: $$ 
2193: Recalling the definition \eqref{eq:Iuv} of  $I_{4,4/3}(5;\m_0)$,
2194: we therefore deduce from the second part of Lemma \ref{lem:B} with $n=11$ that
2195: \begin{align*}
2196:   I_{4,\frac{4}{3}}(5;\m_0) &\ll P^{\frac{49}{4}+2\ve} \left(
2197: \sum_{q}q^{-\frac{5}{6}} \int \min\{1,(|\theta| P^3)^{-\frac{11}{6}}\}
2198:     \d\theta\right)^{\frac{3}{4}},
2199: \end{align*}
2200: where the integral is over $q^{56/31}P^{-3-\delta}\leq |\theta|\leq q^{-44/57}P^{-103/57+\delta}$
2201: and the sum is over $q\leq P^{31/67+2\delta}$,
2202: with the inequalities $q\leq P^\D$ and $|\theta| \leq P^{-3+\D}$ not both
2203: holding simultaneously.
2204: The contribution from $q\leq P^{\D}$ is therefore 
2205: $$
2206: \ll 
2207: P^{10+2\ve-\frac{5\D}{8}} \left(\sum_{q\leq P^\D}q^{-\frac{5}{6}} \right)^{\frac{3}{4}} 
2208: \ll P^{10+2\ve-\frac{\D}{2}},
2209: $$
2210: which is satisfactory.  Taking $|\theta|\geq q^{56/31}P^{-3-\delta}$,
2211: we see that the corresponding contribution from 
2212: $q>P^\D$ is 
2213: $$
2214: \ll 
2215: P^{10+\frac{5\delta}{8}+2\ve}  \left(\sum_{q> P^\D}q^{-\frac{145}{62}} \right)^{\frac{3}{4}} 
2216: \ll P^{10+\frac{5\delta}{8}-\D+2\ve}.
2217: $$
2218: This too is satisfactory, and so completes our analysis of the case $(n_1,n_2)=(2,11)$.
2219: 
2220: 
2221: 
2222: 
2223: 
2224: 
2225: 
2226: 
2227: 
2228: 
2229: 
2230: \subsection{The case $n_1=3$}
2231: 
2232: According to Lemma \ref{lem:curve} we may proceed under the assumption
2233: that either $C_1$ is non-singular or else our cubic form $C$ splits off a
2234: ternary norm form. In the latter case Theorem \ref{main'''} readily
2235: ensures that $X(\QQ)\neq \emptyset$, and so we may focus our efforts
2236: on the case $C_1$ is non-singular. 
2237: We will assume that $w=(w_1,w_2)\in \WW_3^{(1)}\times \WW_{10}^{(2)}$,  throughout this section. 
2238: 
2239: 
2240: Our argument relies upon the same notion of pruning
2241: that was put to good effect in  \S \ref{s:1.12}.
2242: Let us define $\m_1$ to be the set of $\al \in \m$ for which there exists
2243: $a,q\in\ZZ$ such that $0\leq a<q\leq P^{16/25}$ and $\gcd(a,q)=1$,
2244: with 
2245: $$
2246: \Big|\al-\frac{a}{q}\Big|\leq q^{-1}P^{-\frac{143}{75}}.
2247: $$
2248: It will be convenient to refer to the set $\m\setminus\m_1$ as the set of ``proper minor
2249: arcs'', and  $\m_1$ will be the set of ``improper minor
2250: arcs''.  Let us write $E_{\mathrm{prop}}$ and $E_{\mathrm{improp}}$
2251: for the corresponding contributions to $E$.
2252: 
2253: We begin by estimating $E_{\mathrm{prop}}$. 
2254: Taking $u=v=2$ in \eqref{eq:split} and
2255: \eqref{eq:Ivu}, and applying Lemma \ref{lem:C}, we deduce that 
2256: $E_{\mathrm{prop}}\ll I_{2,2}(4;\m\setminus \m_1)$.
2257: When $u=v=2$, $k=4$ and $n_2=10$ we have 
2258: $$
2259:  \rho_{10}=\frac{25}{26}, \quad \pi_{10}=\frac{21}{13},\quad
2260: \rho_{10}''=\frac{7}{11},\quad 
2261: \psi_{10}=\frac{839}{88}=9.53\ldots,
2262: $$
2263: in \eqref{eq:rp1'}, \eqref{eq:r''} and \eqref{eq:psi}.
2264: Now it is easily to see  that $\m\setminus\m_1\subset
2265: \m\setminus\m_0$, where $\m_0$ is as in the statement of Lemma
2266: \ref{lem:v=2}, since for $\al\in \m_0$ we have 
2267: \begin{align*}
2268: \Big|\al-\frac{a}{q}\Big|\leq 
2269: q^{-\frac{1}{3}}P^{-\frac{7}{3}+\delta}=q^{-1}q^{\frac{2}{3}}P^{-\frac{7}{3}+\delta}
2270: &\leq q^{-1}P^{-\frac{21}{11}+2\delta}
2271: <q^{-1}P^{-\frac{143}{75}},
2272: \end{align*}
2273: where $\delta$ is given by \eqref{eq:delta}.
2274: On observing that $\psi_{10}$ satisfies \eqref{eq:in3}, and that
2275: the inequalities in \eqref{eq:in1} are trivially satisfied, it therefore follows from 
2276: Lemma \ref{lem:v=2} that $E_{\mathrm{prop}}=o(P^{10})$.
2277: 
2278: 
2279: We now turn to the argument needed to control the overall contribution
2280: to $E$ from the improper minor arcs, which we denote by
2281: $E_{\mathrm{improp}}$. We select $(A,B,C)=(16/25,1, 82/75)$
2282: in the definition \eqref{eq:Aqa} of $\A=\A(A,B,C)$.
2283: It now follows from taking $n=3$ and $\sigma=-1$ in Lemma \ref{lem:F} that
2284: $$
2285: S_1(\al)-S_1^*(\al)\ll P^{\frac{A}{2}+2+\ve}+P^{2C+\ve}\ll 
2286: P^{\frac{58}{25}+\ve},
2287: $$
2288: for any $\al\in \A_{q,a}$, where $S_1^*(\al)$ is given by \eqref{eq:S*}.
2289: Hence 
2290: $$
2291: E_{\mathrm{improp}}
2292: \ll
2293: \int_{\m_1} |S_1^*(\al)S_2(\al)|\d\al
2294: +P^{\frac{58}{25}+\ve} \int_{\m_1} |S_2(\al)|\d\al=I_1+I_2,
2295: $$
2296: say.  Our goal is to show that $E_{\mathrm{improp}}=o(P^{10})$.
2297: 
2298: We begin by handling the contribution from $I_2$.  
2299: Using dyadic summation it follows from Lemma \ref{lem:B} that
2300: \begin{align*}
2301: I_2
2302: \ll P^{10+\frac{58}{25}+2\ve} \max_{R,\phi} R^{\frac{3}{4}}\phi
2303: \min\{1,(\phi P^3)^{-\frac{5}{4}}\}
2304: &\ll P^{7+\frac{58}{25}+2\ve} R^{\frac{3}{4}}\ll P^{10-\frac{1}{5}+2\ve},
2305: \end{align*}
2306: where the maximum is over $R,\phi$ dictated by the definition of $\m_1$.
2307: This is plainly satisfactory and so completes our treatment of $I_2$.
2308: 
2309: 
2310: 
2311: 
2312: 
2313: We now turn to an upper bound for $I_1$. Combining H\"older's
2314: inequality with the second part of Lemma \ref{lem:F} gives
2315: $$
2316: |I_1|\leq
2317: \left(\int_{\m_1} |S_1^*(\al)|^4\d\al
2318: \right)^{\frac{1}{4}}\left(\int_{\m_1}
2319:   |S_2(\al)|^{\frac{4}{3}}\d\al\right)^{\frac{3}{4}}\ll 
2320: I_{4,\frac{4}{3}}(9;\m_1),
2321: $$
2322: in the notation of \eqref{eq:Iuv}. 
2323: Let us dissect $\m_1$ into $\m_1^{a}\cup\m_1^{b}$, where
2324: $\m_1^{a}$ is the set of $\al\in \m_1$
2325: for which there exist coprime integers $0 \leq a <q$ such that 
2326: $q\leq P^{16/25}$ and  $|\al-a/q|\leq q^{1/2}P^{-3+\delta}$, and 
2327: $\m_1^{b}=\m_1\setminus \m_1^{a}$. An application of Lemma 
2328: \ref{lem:B} yields
2329: \begin{align*}
2330: I_{4,4/3}(9;\m_1^b)
2331: &\ll 
2332: P^{10+\frac{9}{4}+2\ve}\left(
2333: \sum_{q\leq P^{\frac{16}{25}}}q^{-\frac{2}{3}}\int_{q^{\frac{1}{2}}P^{-3+\delta}}^{q^{-1}P^{-\frac{143}{75}}}
2334: (\theta P^3)^{-\frac{5}{3}}\d\theta
2335: \right)^{\frac{3}{4}}\\
2336: &\ll 
2337: P^{10+\frac{9}{4}+2\ve}\left(
2338: \sum_{q}q^{-\frac{2}{3}}P^{-5}
2339: (q^{\frac{1}{2}}P^{-3+\delta})^{-\frac{2}{3}}
2340: \right)^{\frac{3}{4}}\\
2341: &\ll P^{10-\frac{\delta}{2}+2\ve}\log P,
2342: \end{align*}
2343: which is satisfactory.
2344: Turning to $I_{4,4/3}(9;\m_1^a)$, we note that when $(u,v)=(4,4/3)$
2345: and $k=9$ one has
2346: $$
2347:  \rho_{10}=\frac{5}{7}, \quad \pi_{10}=\frac{37}{21},\quad
2348: \rho_{10}'=\frac{8}{7},\quad 
2349: \pi_{10}'=3,
2350: $$ 
2351: in \eqref{eq:rp1} and \eqref{eq:rp2}. Since \eqref{eq:paris}, \eqref{eq:in1} and
2352: \eqref{eq:in2} are evidently satisfied in Lemma~\ref{lem:v<2}, a modest pause for thought reveals that
2353: $I_{4,4/3}(9;\m_1^a)=o(P^{10})$,  as required.
2354: 
2355: 
2356: 
2357: 
2358: 
2359: 
2360: 
2361: \subsection{The case $n_1=4$}
2362: 
2363: 
2364: We follow the strategy of the preceding section. 
2365: According to Lemma \ref{lem:surface} we may assume that either $C_1$ is
2366: non-singular or else the surface $C_1=0$ contains precisely $3$
2367: conjugate double points. In the latter case \eqref{eq:4_norm} implies
2368: that our cubic form $C$ can be written as
2369: $$
2370: \Norm_{K/\QQ}(x_1\omega_1+\cdots+x_3\omega_3)+ax_4^2\Tr_{K/\QQ}(x_1\omega_1+\cdots+x_3\omega_3)+bx_4^3+
2371: C_2(x_5,\ldots,x_{13}),
2372: $$
2373: for appropriate coefficients $\omega_1,\omega_2,\omega_3 \in K$ and
2374: $a,b \in \ZZ$, and where $K$ is a certain cubic number field.
2375: Setting $x_4=0$ we arrive at  a cubic form in $12$ variables which is exactly of the type 
2376: considered in Theorem \ref{main'''}. Hence $X(\QQ)\neq \emptyset$ in
2377: this case, and so  we are free to proceed under the assumption that $C_1$ is non-singular. 
2378: Throughout this section we will take $w=(w_1,w_2)$ as our weight function, with
2379: $w_1\in \WW_4^{(1)}$ and $w_2\in \WW_{9}^{(2)}$.
2380: 
2381: 
2382: Let $\m_1$ be the set of $\al \in \m$ for which there exists
2383: $a,q\in\ZZ$ such that $0\leq a<q\leq P^{21/50}$ and $\gcd(a,q)=1$,
2384: with 
2385: $$
2386: \Big|\al-\frac{a}{q}\Big|\leq q^{-1}P^{-\frac{113}{50}}.
2387: $$
2388: As previously we define $\m\setminus\m_1$ to be the proper minor
2389: arcs and  $\m_1$ to be the improper minor arcs, with the same notation   
2390: $E_{\mathrm{prop}}, E_{\mathrm{improp}}$ for the corresponding
2391: contributions to $E$.
2392: 
2393: We begin by estimating $E_{\mathrm{prop}}$. 
2394: Taking $u=v=2$ in \eqref{eq:split} and
2395: \eqref{eq:Ivu}, and applying Lemma \ref{lem:C}, we find
2396: that $E_{\mathrm{prop}}\ll I_{2,2}(11/2;\m\setminus \m_1)$.
2397: When $u=v=2$, $k=11/2$ and $n_2=9$ we have 
2398: $$
2399:  \rho_{9}=\frac{18}{19}, \quad \pi_{9}=\frac{67}{38},\quad
2400: \rho_{9}''=\frac{25}{62},\quad 
2401: \psi_{9}=9+\frac{15}{124}=9.12\ldots,
2402: $$
2403: in \eqref{eq:rp1'}, \eqref{eq:r''} and \eqref{eq:psi}.
2404: Furthermore, it is easily checked that $\m\setminus\m_1\subset
2405: \m\setminus\m_0$, where $\m_0$ is as in the statement of Lemma
2406: \ref{lem:v=2}.   On observing that \eqref{eq:in3} and 
2407: \eqref{eq:in1} are satisfied, it therefore follows from 
2408: Lemma \ref{lem:v=2} that $E_{\mathrm{prop}}=o(P^{10})$.
2409: 
2410: 
2411: 
2412: It remains to estimate $E_{\mathrm{improp}}$, for which we select
2413: $$
2414: (A,B,C)=\Big(\frac{21}{50},1, \frac{37}{50}\Big)
2415: $$
2416: in the definition \eqref{eq:Aqa} of $\A$.
2417: Taking $n=4$ and $\sigma=-1$ in Lemma \ref{lem:F} therefore gives
2418: $$
2419: S_1(\al)-S_1^*(\al)\ll P^{\frac{5C}{2}+\ve}+P^{\frac{5A}{6}+\frac{5}{2}+\ve}\ll 
2420: P^{\frac{57}{20}+\ve},
2421: $$
2422: for any $\al\in \A_{q,a}$.  It now follows that 
2423: $$
2424: E_{\mathrm{improp}}
2425: \ll
2426: \int_{\m_1} |S_1^*(\al)S_2(\al)|\d\al
2427: +P^{\frac{57}{20}+\ve} \int_{\m_1} |S_2(\al)|\d\al
2428: =I_1+I_2,
2429: $$
2430: say.
2431: We begin by handling the contribution from $I_2$.  
2432: Using dyadic summation it follows from Lemma \ref{lem:B} that
2433: \begin{align*}
2434: I_2
2435: \ll P^{9+\frac{57}{20}+2\ve} \max_{R,\phi} R^{\frac{7}{8}}\phi
2436: \min\{1,(\phi P^3)^{-\frac{9}{8}}\}
2437: &\ll P^{6+\frac{57}{20}+2\ve} R^{\frac{7}{8}}
2438: \ll P^{9.2175+2\ve},
2439: \end{align*}
2440: where the maximum is over $R,\phi$ such that 
2441: $$
2442: 0<R\leq P^{\frac{21}{50}},\quad 0<\phi<R^{-1}P^{-\frac{113}{50}}.
2443: $$
2444: This is plainly satisfactory and so completes our treatment of $I_2$.
2445: 
2446: 
2447: We now turn to an upper bound for $I_1$.  
2448: Since $C_1$ is assumed to be good as well as non-singular, we may
2449: apply the second part of Lemma \ref{lem:F} with $k=3$ and $n=4$ to conclude that
2450: $
2451: I_1\ll I_{3,3/2}(9,\m_1).
2452: $
2453: As in the case $n_1=3$ we write $\m_1=\m_1^{a}\cup\m_1^{b}$, where now
2454: $\m_1^{a}$ is the set of $\al\in \m_1$
2455: for which there exist coprime integers $0 \leq a <q$ such that 
2456: $q\leq P^{21/50}$ and  $|\al-a/q|\leq q P^{-3+\delta}$, and 
2457: $\m_1^{b}=\m_1\setminus \m_1^{a}$. It follows from Lemma 
2458: \ref{lem:B} that 
2459: \begin{align*}
2460: I_{3,\frac{3}{2}}(9;\m_1^b)
2461: &\ll 
2462: P^{12+2\ve}\left(
2463: \sum_{q\leq P^{\frac{21}{50}}}q^{1-\frac{27}{16}}\int_{q P^{-3+\delta}}^{q^{-1}P^{-\frac{113}{50}}}
2464: (\theta P^3)^{-\frac{27}{16}}\d\theta
2465: \right)^{\frac{2}{3}}\\
2466: &\ll 
2467: P^{12+2\ve}\left(
2468: \sum_{q\leq P^{\frac{21}{50}}}q^{-\frac{11}{16}}P^{-\frac{81}{16}}
2469: (q P^{-3+\delta})^{-\frac{11}{16}}
2470: \right)^{\frac{2}{3}}\\
2471: &\ll P^{10-\frac{11\delta}{24}+2\ve}.
2472: \end{align*}
2473: To handle $I_{3,3/2}(9;\m_1^a)$ we note that when $(u,v)=(3,3/2)$
2474: and $k=9$ we have 
2475: $$
2476:  \rho_{9}=\frac{3}{4}, \quad \pi_{9}=\frac{29}{16},\quad
2477: \rho_{9}'=\frac{43}{25},\quad 
2478: \pi_{9}'=3,
2479: $$ 
2480: in \eqref{eq:rp1} and \eqref{eq:rp2}. 
2481: Lemma~\ref{lem:v<2} easily gives 
2482: $I_{3,3/2}(9;\m_1^a)=o(P^{10})$,  as required.
2483: This completes the treatment of the improper minor arcs when $(n_1,n_2)=(4,9)$.
2484: 
2485: 
2486: 
2487: 
2488: 
2489: 
2490: 
2491: \subsection{The case $n_1=5$}
2492: 
2493: An application of Lemma \ref{lem:3fold} reveals that we are free to assume
2494: that $C_1$ defines a projective cubic hypersurface whose singular locus is either empty or finite.
2495: Throughout this section we will take $w=(w_1,w_2)\in \WW_5^{(1)}\times
2496: \WW_{8}^{(2)}$ as our weight function.
2497: 
2498: We let the improper minor arcs $\m_1$ be the set of $\al \in \m$ for which there exists
2499: $a,q\in\ZZ$ such that $0\leq a<q\leq P^{\frac{11}{20}+\delta}$ and $\gcd(a,q)=1$,
2500: with 
2501: $$
2502: \Big|\al-\frac{a}{q}\Big|\leq P^{-\frac{5}{2}+\delta},
2503: $$
2504: with $\delta$ given by \eqref{eq:delta}, and we let 
2505: $\m\setminus\m_1$  be the proper minor
2506: arcs.  As above, let $E_{\mathrm{prop}}, E_{\mathrm{improp}}$ denote
2507: the corresponding contributions to $E$.
2508: 
2509: 
2510: Taking $u=v=2$ in \eqref{eq:split} and
2511: \eqref{eq:Ivu}, and applying Lemma \ref{lem:C} with $n=5$ and
2512: $\sigma\leq 0$, we find
2513: that $E_{\mathrm{prop}}\ll I_{2,2}(15/2;\m\setminus \m_1)$.
2514: When $u=v=2$, $k=15/2$ and $n_2=8$ we have 
2515: $$
2516:  \rho_{8}=\frac{12}{13}, \quad \pi_{8}=\frac{22}{13},\quad
2517: \rho_{8}''=\frac{11}{20},\quad 
2518: \psi_{8}=9.55,
2519: $$
2520: in \eqref{eq:rp1'}, \eqref{eq:r''} and \eqref{eq:psi}.
2521: Furthermore, it is easily checked that $\m\setminus\m_1\subset
2522: \m\setminus\m_0$, where $\m_0$ is as in the statement of Lemma
2523: \ref{lem:v=2}.   On observing that \eqref{eq:in1} and \eqref{eq:in3}
2524: are satisfied, it therefore follows from 
2525: Lemma \ref{lem:v=2} that $E_{\mathrm{prop}}=o(P^{10})$.
2526: 
2527: 
2528: 
2529: It remains to estimate $E_{\mathrm{improp}}$, for which we select
2530: $$
2531: (A,B,C)=\Big(\frac{11}{20}+\delta,0, \frac{1}{2}+\delta\Big)
2532: $$
2533: in the definition \eqref{eq:Aqa} of $\A$.
2534: Taking $n=5$ and $\delta\leq 0$ in Lemma \ref{lem:F} therefore gives
2535: $$
2536: S_1(\al)-S_1^*(\al)\ll 
2537: P^{3+\frac{5A}{3}+\ve}
2538: +P^{3(A+C)+\ve}\ll P^{3.92},
2539: $$
2540: for any $\al\in \A_{q,a}$.  It now follows that 
2541: $$
2542: E_{\mathrm{improp}}
2543: \ll
2544: \int_{\m_1} |S_1^*(\al)S_2(\al)|\d\al
2545: +P^{3.92} \int_{\m_1} |S_2(\al)|\d\al
2546: =I_1+I_2,
2547: $$
2548: say.
2549: 
2550:    
2551: We begin by handling the contribution from $I_2$.  
2552: Using dyadic summation it follows from Lemma \ref{lem:B} that
2553: \begin{align*}
2554: I_2
2555: \ll P^{3.92+\ve} \max_{R,\phi} R \phi
2556: P^{8}\min\{1,(\phi P^3)^{-1}\}
2557: \ll P^{8.92+\ve} R\ll P^{9.47+\delta +\ve},
2558: \end{align*}
2559: where the maximum is over $R,\phi$ such that 
2560: $$
2561: 0<R\leq P^{\frac{11}{20}+\delta},\quad 0<\phi<P^{-\frac{5}{2}+\delta}.
2562: $$
2563: This is plainly satisfactory and so completes our treatment of $I_2$.
2564: 
2565: 
2566: We now turn to an upper bound for $I_1$. 
2567: Since $C_1$ is assumed to be good, we may
2568: apply the second part of Lemma \ref{lem:F} with $k=12/5$ and $n=5$ to conclude that
2569: $I_1\ll I_{12/5,12/7}(9;\m_1)$, in the notation of \eqref{eq:Iuv}.
2570: When $(u,v)=(12/5,12/7)$, $k=9$ and $n_2=8$ we have 
2571: $$
2572:  \rho_{8}=\frac{4}{5}, \quad \pi_{8}=\frac{66}{35},\quad
2573: \rho_{8}'=\frac{38}{11},\quad 
2574: \pi_{8}'=3,
2575: $$
2576: in \eqref{eq:rp1} and \eqref{eq:rp2}.
2577: Furthermore one easily checks that $k/u=15/4$ satisfies the
2578: inequalities in \eqref{eq:in1}. It therefore follows from
2579: Lemma \ref{lem:v<2} that $I_{12/5,12/7}(9;\m_1\setminus \m_2)=o(P^{10})$, where
2580: $\m_2$ is the set of $\al\in \m_1$ for which there exist coprime
2581: integers $0\leq a<q$ such that
2582: \begin{equation}
2583:   \label{eq:cut}
2584: q\leq P^{\frac{11}{42}+2\delta},\quad q^{\frac{38}{11}}P^{-3-\delta}\leq
2585: \Big|\al-\frac{a}{q}\Big|\leq q^{-\frac{4}{5}}P^{-\frac{66}{35}+\delta}. 
2586: \end{equation}
2587: Note that $q^{-4/5}P^{-66/35+\delta}\leq q^{-1}P^{-3/2}$. 
2588: To handle $I_{12/5,12/7}(9;\m_2)$ we appeal to Lemma
2589: \ref{lem:B}, deducing that 
2590: \begin{align*}
2591: I_{\frac{12}{5},\frac{12}{7}}(9;\m_2)
2592: &\ll P^{8+\frac{15}{4}+2\ve} \max_{R,\phi} (R^2 \phi)^{\frac{7}{12}}
2593: R^{-1}\min\{1,(\phi P^3)^{-1}\}\\
2594: &\ll 
2595: \max_{R,\phi} P^{5+\frac{15}{4}+2\ve}
2596: R^{\frac{1}{6}}\phi^{-\frac{5}{12}}\\
2597: &\ll P^{10+\frac{\delta}{2}+2\ve}R^{-\frac{14}{11}},
2598: \end{align*}
2599: on taking $\phi\geq R^{38/11}P^{-3-\delta}$.
2600: Here the maximum is over the relevant $R,\phi$ determined by \eqref{eq:cut},
2601: with the inequalities $R\leq P^\D$ and $\phi \leq P^{-3+\D}$ not both
2602: holding simultaneously.
2603: Now either $R>P^\D$ and this estimate is satisfactory, or else $R\leq
2604: P^{\D}$ and it follows that we may actually take the lower bound
2605: $\phi>P^{-3+\D}$ in the second term, giving instead
2606: $$
2607: \ll P^{10-\frac{5\D}{12}+2\ve} R^{\frac{1}{6}}\ll P^{10-\frac{\D}{4}+2\ve}.
2608: $$
2609: This too is satisfactory, and so completes the proof that
2610: $I_1=o(P^{10})$, thereby completing the treatment of the minor arcs in
2611: the case $(n_1,n_2)=(5,8)$.
2612: 
2613: 
2614: 
2615: 
2616: 
2617: \subsection{The case $n_1=6$}
2618: 
2619: We now come to the final case that we need to analyse
2620: in our proof of Theorem \ref{main}.
2621: We take $w=(w_1,w_2)\in  \WW_6^{(2)}\times \WW_{7}^{(2)}$, and seek 
2622: an estimate for
2623: $$
2624:   M_6(P;\m):=\int_{\m}|S_1(\al)|^{2}\d\al.
2625: $$
2626: Note that the integral is now taken over the set of minor arcs, rather
2627: than the entire interval $[0,1]$ as in \eqref{eq:moment}.
2628: As usual we assume that $C_1$ and $C_2$ are good. 
2629: We will show that 
2630: \begin{equation}
2631:   \label{eq:log}
2632:  M_6(P;\m)\ll P^{9+\ve}. 
2633: \end{equation}
2634: Once in place, we may take 
2635: $u=v=2$ in \eqref{eq:split} and \eqref{eq:Iuv} to conclude that
2636: $
2637: E\ll I_{2,2}(9;\m),
2638: $
2639: whence the desired conclusion is
2640: given by Lemma \ref{lem:v=2'}.
2641: 
2642: 
2643: It remains to establish \eqref{eq:log}. 
2644: Let us consider the consequences of applying Lemma \ref{lem:D} to
2645: estimate $M=M_6(P;\m)$, following the general approach in the proof of
2646: Lemmas \ref{lem:v<2} and \ref{lem:v=2}.
2647: We have 
2648: $$
2649: M
2650: \ll P^3+P^\ve \max_{R,\phi}
2651: \frac{\psi_H R^2 P^{11}}{H^{10}}F,
2652: $$
2653: where $\psi_H$ and $F$ are as in the statement of Lemma \ref{lem:D}
2654: and $H\in [1,P]\cap \ZZ$ is arbitrary. Furthermore the maximum is over
2655: $R,\phi$ such that
2656: \eqref{eq:range} holds with any choice of $Q\geq 1$ that we care
2657: to choose. We will take $Q=P^{3/2}$.
2658: In our deduction of \eqref{eq:log} it will be convenient to allow the value of $\ve>0$
2659: to take different values at different parts of the argument. 
2660: 
2661: We define $\phi_0:=R^{-2/11}P^{-2}$ and take
2662: $$
2663: H:=
2664: \begin{cases}
2665: \lfloor P^{\ve}\max\{1,(\phi R^{2}P^{2})^{\frac{1}{10}}\} \rfloor, &\mbox{if  $\phi>\phi_0$,}\\
2666: \lfloor P^{\ve} R^{\frac{2}{11}} \rfloor, &\mbox{if  $\phi\leq \phi_0$.}
2667: \end{cases}
2668: $$
2669: If we can show that $F\ll P^{\ve}$ with this
2670: choice of $H$ then \eqref{eq:log} will follow.  It is clear that $H$ is an integer in the
2671: interval $[1,P]$.
2672: 
2673: Suppose first that $\phi>\phi_0$. Then $\psi_H\ll \phi$ and it follows that
2674: \begin{align*}
2675: RH^3\psi_H
2676: \ll P^\ve R\phi (1+\phi R^2P^2)^{\frac{3}{10}}
2677: &\ll 1+ Q^{-1}P^{\frac{3}{5}+\ve}\ll 1,
2678: \end{align*}
2679: by \eqref{eq:range} and the fact that $Q=P^{3/2}$.
2680: The third term in $F$ is 
2681: \begin{align*}
2682: \frac{H^{6}}{R^{3}(P^2\psi_H)^{2}}\ll 
2683: \frac{P^{\ve}}{R^{3}P^{4}\phi^{2}}
2684: (1+\phi R^2P^2)^{\frac{3}{5}}
2685: &\ll
2686: \frac{P^{\ve}}{R^{3}P^{4}\phi_0^{2}}
2687: +\frac{P^{\ve}}{R^{\frac{9}{5}}P^{\frac{14}{5}}\phi_0^{\frac{7}{5}}}
2688: \ll P^{\ve},
2689: \end{align*}
2690: which is satisfactory. 
2691: 
2692: Suppose now that $\phi\leq \phi_0$.  Then $\psi_H\ll (P^2H)^{-1}$ and it follows that
2693: \begin{align*}
2694: RH^3\psi_H
2695: &\ll \frac{RH^2}{P^2}\ll \frac{RP^\ve}{P^2}(1+R^{\frac{2}{11}})\ll
2696: 1+\frac{Q^{\frac{13}{11}}P^\ve}{P^2}\ll 1.
2697: \end{align*}
2698: Turning to the third term in $F$, we find that
2699: \begin{align*}
2700: \frac{H^{6}}{R^{3}(P^2\psi_H)^{2}}\ll 
2701: \frac{H^{8}}{R^{3}}
2702: \ll
2703: \frac{P^{\ve}}{R^{3}}
2704: (1+R^{\frac{16}{11}})
2705: \ll P^\ve,
2706: \end{align*}
2707: which is also satisfactory. This therefore completes the proof of
2708: \eqref{eq:log}, and so our treatment of the case $(n_1,n_2)=(6,7)$.
2709: 
2710: 
2711: 
2712: 
2713: 
2714: 
2715: 
2716: 
2717: 
2718: 
2719: 
2720: 
2721: 
2722: 
2723: 
2724: 
2725: 
2726: 
2727: 
2728: 
2729: 
2730: 
2731: 
2732: 
2733: 
2734: 
2735: 
2736: 
2737: \begin{thebibliography}{99}
2738: 
2739: \bibitem{baker} 
2740: R.C. Baker, 
2741: Diagonal cubic equations, II, {\em Acta Arith.} {\bf 53} (1989),  217--250.
2742: 
2743: 
2744: \bibitem{BDL}
2745: B.J. Birch, H. Davenport and D.J. Lewis,
2746: The addition of norm forms. {\em Mathematika} {\bf 9} (1962), 75--82.
2747: 
2748: 
2749: \bibitem{cubhyp-circle}
2750: T.D. Browning,
2751: Counting rational points on cubic hypersurfaces.
2752: {\em Mathematika} {\bf 54} (2007), 93--112.
2753: 
2754: 
2755: \bibitem{roth} 
2756: T.D. Browning and D.R. Heath-Brown,
2757: Integral points on cubic hypersurfaces.
2758: {\em Analytic Number Theory: Essays in honour of Klaus Roth}, 
2759: Cambridge University Press, to appear.
2760: 
2761: 
2762: \bibitem{41}
2763: T.D. Browning and D.R. Heath-Brown, Rational points on quartic
2764: hypersurfaces. {\em J. reine angew. Math.}, to appear.
2765: 
2766: %\bibitem{brud}
2767: %J. Br\"udern,
2768: %Ternary additive problems of Waring's type.
2769: %{\em Math. Scand.} {\bf 68} (1991), 27--45.
2770: 
2771: \bibitem{b-w}
2772: {J.W. Bruce   and C.T.C. Wall},
2773: {On the classification of cubic surfaces}.
2774: {\em J. London Math. Soc.} {\bf 19} (1979), {245--256}.
2775: 
2776: 
2777: 
2778: \bibitem{cayley}
2779: A. Cayley, A memoir on cubic surfaces. {\em Phil. Trans. Roy. Soc.}
2780: {\bf 159} (1869), 231--326.
2781: 
2782: \bibitem{c-t-s}
2783: J.-L. Colliot-Th\'el\`ene and P. Salberger,
2784: Arithmetic on some singular cubic hypersurfaces.
2785: {\em Proc. London Math. Soc.}  {\bf 58} (1989),  519--549.
2786: 
2787: 
2788: \bibitem{ct2}
2789: J.-L. Colliot-Th\'el\`ene, J.-J. Sansuc and P. Swinnerton-Dyer, 
2790: Intersections of two quadrics and Ch\^atelet surfaces. II.  
2791: {\em J. reine angew. Math.}  {\bf 374}  (1987), 72--168. 
2792: 
2793: 
2794: %\bibitem{coray76a}
2795: %D. F. Coray,
2796: %Arithmetic on singular cubic surfaces.
2797: %{\em Compositio Math.}  {\bf 33} (1976), 55--67.
2798: 
2799: 
2800: \bibitem{coray76b}
2801: D. F. Coray,
2802: Algebraic points on cubic hypersurfaces.
2803: {\em Acta Arith.}  {\bf 30} (1976), 267--296.
2804: 
2805: \bibitem{coray87}
2806: D. F. Coray,
2807: Cubic hypersurfaces and a result of Hermite.
2808: {\em Duke Math. J.}  {\bf 54} (1987), 657--670.
2809: 
2810: 
2811: \bibitem{c-t}
2812: D. F. Coray and M. A. Tsfasman,
2813: Arithmetic on singular Del Pezzo surfaces.
2814: {\em Proc. London Math. Soc.}  {\bf 57} (1988),  25--87.
2815: 
2816: 
2817: \bibitem{6}
2818: D. F. Coray,
2819: D.J. Lewis, N. Shepherd-Barron and P. Swinnerton-Dyer,
2820: Cubic threefolds with six double points.  {\em 
2821: Number theory in progress, Vol. 1 (Zakopane-Ko\'scielisko, 1997)},
2822: 63--74, de Gruyter, Berlin, 1999. 
2823: 
2824: 
2825: \bibitem{dav-32} 
2826: H. Davenport, 
2827: Cubic forms in thirty-two variables.
2828: {\em Philos. Trans. Roy. Soc. London. Ser. A} {\bf 251} (1959), 193--232. 
2829: 
2830: 
2831: \bibitem{dav-16} 
2832: H. Davenport, 
2833: Cubic forms in sixteen variables.
2834: {\em Philos. Trans. Roy. Soc. London. Ser. A} {\bf 272} (1963), 285--303. 
2835: 
2836: 
2837: \bibitem{dav-book}
2838: H. Davenport, {\em Analytic methods for Diophantine 
2839: equations and Diophantine inequalities.} 2nd ed., CUP, 2005.
2840: 
2841: \bibitem{fowler} J. Fowler, 
2842: A note on cubic equations. {\em Proc. Camb. Philos. Soc.} {\bf 58} 
2843: (1962), 165--169.
2844: 
2845: \bibitem{f-m-t}
2846: J. Franke, Y.I. Manin and Y. Tschinkel,
2847: Rational points of bounded height on {F}ano varieties.
2848: {\em Invent. Math.} {\bf 95} (1989), 421--435.
2849: 
2850: \bibitem{harris} 
2851: J. Harris, {\em Algebraic Geometry}, Springer-Verlag, 1992.
2852: 
2853: 
2854: 
2855: \bibitem{hb-10} 
2856: D.R. Heath-Brown,
2857: Cubic forms in ten variables.
2858: {\em Proc. London. Math. Soc.}
2859: {\bf 47} (1983),  225--257.
2860: 
2861: \bibitem{14}D.R. Heath-Brown, Rational points on cubic
2862: hypersurfaces. {\em Invent. Math.} {\bf 170} (2007), 199--230.
2863: 
2864: \bibitem{hooley1}
2865: C. Hooley, On nonary cubic forms.  {\em J. reine angew. Math.}
2866: {\bf 386} (1988), 32--98.
2867: 
2868: 
2869: \bibitem{hooley2}
2870: C. Hooley, On nonary cubic forms. II. {\em J. reine angew. Math.}
2871: {\bf 415} (1991), 95--165.
2872: 
2873: %\bibitem{hooley-38}
2874: %C. Hooley, On the number of points on a complete intersection over a
2875: %finite field.{\em J. Number Theory} {\bf 38} (1991), 338--358.
2876: 
2877: \bibitem{kollar}
2878: J. Koll\'ar,
2879: Unirationality of cubic hypersurfaces.  
2880: {\em J. Inst. Math. Jussieu}  {\bf 1}  (2002),  467--476. 
2881: 
2882: 
2883: \bibitem{lewis}
2884: D.J. Lewis,
2885: Diophantine problems: solved and unsolved.
2886: {\em Number theory and applications}, 103--121, Kluwer, Dordrecht, 1989.
2887: 
2888: \bibitem{manin}
2889: Y.I. Manin, {\em Cubic forms}. 2nd ed.,
2890: North-Holland Mathematical Library {\bf 4}, North-Holland
2891: Publishing Co.,  1986. 
2892: 
2893: \bibitem{cayley'}
2894: L. Schl\"afli, On the distribution of surfaces of the third order into
2895: species. {\em Phil. Trans. Roy. Soc.}
2896: {\bf 153} (1864), 193--247.
2897: 
2898: \bibitem{segre-b}
2899: B. Segre,  On arithmetical properties of singular cubic surfaces.
2900: {\em  J. London Math. Soc.} {\bf 19} (1944), 84--91.
2901: 
2902: \bibitem{segre}
2903: C. Segre, Suella varieta cubica con dieci punti doppi dello spazio a
2904: quattro dimensioni.
2905: {\em Atta della R. Accademia della Scienae di Torino} {\bf 22}
2906: (1886--87), 791--801.
2907: 
2908: \bibitem{skolem}
2909: T. Skolem, 
2910: Einige Bemerkungen \"uber die Auffindung der rationalen Punkte auf
2911: gewissen algebraischen Gebilden. {\em Math. Zeit.} {\bf 63}  (1955),
2912: 295--312.  
2913: 
2914: \bibitem{vaughan}
2915: R.C. Vaughan, 
2916: {\em The Hardy--Littlewood method}, 2nd ed., CUP, 1997.
2917: 
2918: 
2919: \bibitem{wooley}
2920: T. Wooley,
2921: On Weyl's inequality, Hua's lemma and exponential sums over binary forms.
2922: {\em Duke Math. J.} {\bf 100} (1999), 373--423.
2923: 
2924: 
2925: \end{thebibliography} 
2926: 
2927: 
2928: 
2929: \newpage
2930: \appendix
2931: 
2932:  \setcounter{section}{-1}
2933: \section{Appendice par J.-L. Colliot-Th\'el\`ene}
2934: 
2935: \begin{center}
2936: \textsc{Groupe de Brauer non ramifi\'e des hypersurfaces
2937: cubiques\\  singuli\`eres  (d'apr\`es P. Salberger)}
2938: \end{center}
2939: 
2940: 
2941: \bigskip
2942: 
2943: 
2944: 
2945: En r\'eponse \`a une question de R. Heath-Brown, P. Salberger en 2006
2946: a indiqu\'e les grandes lignes de la d\'emonstration de l'\'enonc\'e suivant, qui
2947: \'etend un r\'esultat
2948: connu  dans le cas lisse (\cite{CT}).   Nous donnons le d\'etail de la d\'emonstration.
2949: On utilise les notations usuelles
2950: dans ce domaine. Pour $X$ un sch\'ema on note $\pic X= H^1_{\et}(X,\G_{m})$
2951: son groupe de Picard et $\br X=H^2_{\et}(X,\G_{m})$ son groupe de Brauer.
2952: Pour les propri\'et\'es usuelles de ces groupes, nous renvoyons le
2953: lecteur \`a   \cite{CTSan}.
2954: 
2955: 
2956: \begin{theorem-f}
2957: Soit $k$ un corps de caract\'eristique z\'ero, $\kbar$ une cl\^oture alg\'ebrique,
2958: $\g$ le groupe de Galois de $\kbar$ sur $k$.  Soit $X \subset {\bf P}^n_{k}$
2959: une  intersection compl\`ete  g\'eo\-m\'etriquement int\`egre de dimension au moins $3$.
2960: Supposons le lieu singulier vide
2961: ou de codimension au moins \'egale \`a $4$ dans $X$.
2962: Alors pour tout   $k$-mod\`ele projectif et   lisse $Y$ de $X$,
2963: \begin{itemize}
2964: \item[(a)]
2965: le groupe de Picard de $\Y=Y\times_{k}\kbar$ est un module galoisien $\Z$-libre de type
2966: fini
2967: qui est stablement de pemutation;
2968: \item[(b)]
2969: on a $H^1(\g,\pic \Y)=0$;
2970: \item[(c)]
2971: on a $\br k    {\buildrel \simeq \over \rightarrow}   {\rm Ker} [ \br Y \to \br \Y]     .$
2972: \end{itemize}
2973: \end{theorem-f}
2974: 
2975: 
2976: \begin{proof}[D\'emonstration]
2977: Les anneaux locaux de $X$ en codimension au
2978: moins 3 sont r\'e\-gu\-liers, donc factoriels (th\'eor\`eme
2979: d'Auslan\-der-Buchsbaum, voir \cite[\S XI Thm. 3.13]{SGA 2}).
2980: Comme $X$ est une intersection compl\`ete,
2981: un  th\'eor\`eme de Grothendieck  \cite[\S XI Cor. 3.14]{SGA 2}, ex-conjecture de Samuel)
2982: implique que tous les anneaux locaux de $X$ sont factoriels.
2983: Ainsi les diviseurs de Weil sur $X$ sont tous des diviseurs de Cartier.
2984: Ceci implique que pour tout ouvert $U \subset X$ la fl\`eche de restriction $\pic X \to
2985: \pic U$ est surjective.
2986: Cette fl\`eche est aussi injective. Soit en effet $D$ un diviseur sur $X$
2987: qui est le diviseur d'une fonction rationnelle $f$ sur $U$.
2988: Comme le compl\'ementaire de $U$ dans $X$ est de codimension
2989: au moins 2 et que sur $X$ diviseurs de Weil et diviseurs de Cartier
2990: co\"{\i}ncident, on conclut que $D$ est le diviseur de $f$ sur $X$.
2991: Le m\^eme argument montre que toute fonction rationnelle sur $X$
2992: d\'efinie et  inversible sur $U$ est d\'efinie et inversible sur $X$,
2993: et comme $X$ est projectif et g\'eom\'etriquement int\`egre,
2994: toute telle fonction est une constante, elle appartient \`a  $k^*$.
2995: 
2996: L'hypoth\`ese sur la codimension du lieu singulier est g\'eom\'etrique,
2997: elle vaut pour $X_{K}$ pour toute extension $K/k$ de corps, par
2998: exemple $\kbar/k$. Les m\^emes conclusions s'appliquent
2999: donc \`a $\X$. En particulier
3000: la fl\`eche de restriction
3001: $\pic \X \to \pic \U $ est   un isomorphisme.
3002: 
3003: Par ailleurs, le Corollaire 3.7 de \cite[\S XII]{SGA 2} montre que la fl\`eche de
3004: restriction
3005: $$
3006: \Z= \pic  {\bf P}^n_{k} \to \pic X
3007: $$
3008: qui envoie la classe de $1 \in \Z$
3009: sur la classe du faisceau inversible ${\mathcal O}_{X}(1)$
3010: est un isomorphisme. Il en est de m\^eme de
3011:   $\Z= \pic  {\bf P}^n_{\kbar} \to \pic  \X $,
3012: et l'action du groupe
3013: de Galois sur $\Z    \simeq  \pic  \X $ est triviale.
3014: 
3015: En conclusion, sous les hypoth\`eses du th\'eor\`eme, le module galoisien
3016: $\pic \U$ est le module galoisien trivial $\Z$   et
3017: l'on a $\kbar^*   {\buildrel \simeq \over \rightarrow}   \kbar[U]^*$, o\`u $\kbar[U]$ est l'anneau des
3018: fonctions d\'efinies sur $\U$. L'argument ci-dessus
3019: montre aussi que l'application naturelle
3020: $\pic U \to \pic \U$ est un isomorphisme.
3021: 
3022: D'une suite exacte bien connue (cf. \cite[p.386]{CTSan})
3023: on d\'eduit que la fl\`eche naturelle
3024: $\br k \to {\rm Ker} [\br U \to \br \U]$ est un isomorphisme.
3025: Par des arguments standards sur le groupe de Brauer (puret\'e et
3026: injection par passage d'une vari\'et\'e lisse \`a un ouvert)
3027: % d\'etailler
3028: un tel \'enonc\'e
3029: implique le m\^eme \'enonc\'e pour toute $k$-vari\'et\'e projective et
3030: lisse $k$-birationnelle \`a $X$ : c'est l'\'enonc\'e (c).
3031: 
3032: 
3033: 
3034: Soit $U \subset Y$ une $k$-compactification lisse de $U$
3035: (le th\'eor\`eme de Hironaka assure l'existence d'une telle compactification).
3036: On a alors la suite exacte de modules galoisiens
3037: $$0 \to \Div_{\infty}\Y \to \pic \Y \to \pic \U  \to 0,$$
3038: o\`u le groupe $\Div_{\infty}\Y $ est
3039: le module de permutation sur les points de codimension 1 de $\Y $ en dehors de $\U $,
3040: et le z\'ero
3041: \`a gauche tient au fait que l'on a $\kbar^* \simeq \kbar[U]^*$.
3042: La suite de modules galoisiens ci-dessus est scind\'ee, car tout groupe
3043: $H^1(\g, P)$ \`a valeurs dans un module de permutation est nul (lemme de Shapiro).
3044: Ainsi $\pic \Y$ est la somme directe de deux modules de permutation,
3045: et est donc un module de permutation. Il en r\'esulte que pour tout autre mod\`ele
3046: projectif
3047: et lisse $Y'$, le module galoisien $\pic  \Y'$ est stablement de permutation
3048: (\cite[Prop. 2.A.1 on p. 461]{CTSan}).
3049: L'\'enonc\'e (b) en r\'esulte.
3050: \end{proof}
3051: 
3052: 
3053: 
3054: 
3055: \begin{cor-f}
3056: Soit $X \subset {\bf P}^n_{k}$ une hypersurface cubique g\'eom\'etriquement in\-t\`egre
3057: de dimension au moins $3$ qui n'est pas un c\^one.
3058: Supposons le lieu singulier vide
3059: ou de codimension au moins \'egale \`a $4$ dans $X$. Alors pour tout mod\`ele projectif
3060: et lisse
3061: $Y$ de $X$, on a $\br k  {\buildrel \simeq \over \rightarrow}  \br Y$.
3062: \end{cor-f}
3063: 
3064: \begin{proof}[D\'emonstration] Au vu du th\'eor\`eme ci-dessus, il suffit de montrer
3065: $\br \Y =0$.
3066: 
3067: Si l'hypersurface $X$ est lisse, on a $\br \X =0$  comme il est
3068: \'etabli dans \cite{CT} sans restriction sur le degr\'e de
3069: l'hypersurface.
3070: Par l'invariance birationnelle du groupe de
3071: Brauer pour les vari\'et\'es projectives et lisses ceci implique $\br \Y =0$.
3072: 
3073: Si  l'hypersurface cubique $X$ est singuli\`ere,
3074: comme elle n'est pas un c\^one,
3075: en utili\-sant les droites passant par un $\kbar$-point singulier on obtient
3076: une \'equivalence birationnelle  de $\X$
3077: avec l'espace projectif ${\bf P}^{n-1}_{\kbar}$, dont le groupe de Brauer est nul.
3078: Par l'invariance birationnelle du groupe de Brauer
3079: pour les vari\'et\'es projectives et lisses ceci implique $\br \Y =0$.
3080: \end{proof}
3081: 
3082: 
3083: \begin{remark}
3084: Il serait int\'eressant de voir si le corollaire vaut pour
3085: les hypersurfaces   de degr\'e sup\'erieur \`a 3.
3086: C'est le cas lorsque les hypersurfaces sont lisses (\cite{CT}).
3087: \end{remark}
3088: 
3089: \begin{remark}
3090: La condition que la codimension du lieu singulier est au moins \'egale \`a 4
3091: est n\'ecessaire. Dans \cite{CTSal} on trouve des hypersurfaces
3092: cubiques g\'eom\'etrique\-ment int\`egres non coniques dans ${\bf P}^4_{k}$,
3093: dont le lieu singulier est  un ensemble fini non vide de points,
3094: et  qui admettent un mod\`ele projectif et lisse $Y$
3095: avec $\br Y/\br k \neq 0$.
3096: \end{remark}
3097: 
3098: \begin{remark}
3099: Lorsque $k$ est un corps de nombres, le corollaire permet de conjecturer
3100: que, sous les hypoth\`eses donn\'ees, le principe de Hasse et l'approxi\-mation faible
3101: valent pour le lieu lisse de l'hypersurface cubique $X$.
3102: \end{remark}
3103: 
3104: 
3105: 
3106: 
3107: 
3108: \renewcommand{\refname}{R\'ef\'erences}
3109: \begin{thebibliography}{99}
3110: 
3111: 
3112: \bibitem{CT}
3113: J.-L. Colliot-Th\'el\`ene,
3114: The Brauer-Manin obstruction for complete intersections
3115: of dimension $\geq 3$, appendix to \cite{PV}.
3116: 
3117: \bibitem{CTSal}
3118: J.-L. Colliot-Th\'el\`ene and P. Salberger,
3119: Arithmetic on some singular cubic hypersurfaces.
3120: {\em Proc. London Math. Soc.}  {\bf 58} (1989),  519--549.
3121: 
3122: \bibitem{CTSan}
3123: J.-L. Colliot-Th\'el\`ene et J.-J. Sansuc,
3124: La descente sur les vari\'et\'es rationnelles, II,
3125: {\em Duke Math. J.} {\bf 54} (1987) 375--492.
3126: 
3127: \bibitem{SGA 2}
3128: A. Grothendieck,
3129: Cohomologie locale des faisceaux coh\'erents et th\'eor\`emes de
3130: Lefschetz locaux et globaux (SGA 2). {\em S\'eminaire de G\'eom\'etrie
3131: Alg\'ebrique du Bois Marie, 1962. Augment\'e d'un expos\'e de
3132: Mich\`ele Raynaud}, Documents Math\'ematiques (Paris) {\bf
3133:   4}, Soci\'et\'e Math\'ematique de France, 2005.
3134: 
3135: 
3136: \bibitem{PV}
3137: B. Poonen and J.F. Voloch, Random Diophantine equations.
3138: {\em Arithmetic of higher-dimensional algebraic varieties (Palo Alto,
3139: CA, 2002)}, 175--184, Progr. Math. {\bf 226}, Birkh\"auser, 2004.
3140: 
3141: 
3142: \end{thebibliography} 
3143: 
3144: 
3145: 
3146: 
3147: 
3148: \begin{flushright}
3149: \scriptsize{
3150: J.-L. Colliot-Th\'el\`ene}\\
3151: \scriptsize{C.N.R.S., UMR 8628 CNRS-Universit\'e}\\
3152: \scriptsize{Math\' ematiques, B\^atiment 425}\\
3153: \scriptsize{Universit\'e   Paris-Sud}\\
3154: \scriptsize{F-91405 Orsay}\\
3155: \scriptsize{France}\\
3156: \texttt{jlct@math.u-psud.fr}
3157: \end{flushright}
3158: 
3159: 
3160: 
3161: 
3162: \end{document}
3163: 
3164: 
3165: