1: \documentstyle[fleqn,twoside]{article}
2:
3: \makeatletter
4:
5: %% PAGE SETUP
6:
7: \topmargin -8mm
8: \oddsidemargin -7mm
9: \evensidemargin -12mm
10: \textheight 240mm
11: \textwidth 176mm
12: \columnsep 8mm
13: \columnseprule 0.2pt
14: \emergencystretch 6pt
15: \mathsurround 1pt
16: \mathindent 1em
17: \pagestyle{myheadings}
18: \newcommand{\bls}[1]{\renewcommand{\baselinestretch}{#1}}
19: \def\onecol{\onecolumn \mathindent 2em}
20: \def\twocol{\twocolumn \mathindent 1em}
21: \def\noi{\noindent}
22:
23: %% SECTIONING AND HEADINGS
24:
25: \renewcommand{\section}{\@startsection{section}{1}{0pt}%
26: {-3.5ex plus -1ex minus -.2ex}{2.3ex plus .2ex}%
27: {\large\bf\protect\raggedright}}
28: \renewcommand{\thesection}{\arabic{section}.}
29: \renewcommand{\subsection}{\@startsection{subsection}{2}{0pt}%
30: {-3ex plus -1ex minus -.2ex}{1.4ex plus .2ex}%
31: {\normalsize\bf\protect\raggedright}}
32: \renewcommand{\thesubsection}{\arabic{section}.\arabic{subsection}.}
33: \renewcommand{\thesubsubsection}%
34: {\arabic{section}.\arabic{subsection}.\arabic{subsubsection}.}
35:
36: \renewcommand{\@oddhead}{\raisebox{0pt}[\headheight][0pt]{%
37: \vbox{\hbox to\textwidth{\rightmark \hfil \rm \thepage \strut}\hrule}}}
38: \renewcommand{\@evenhead}{\raisebox{0pt}[\headheight][0pt]{%
39: \vbox{\hbox to\textwidth{\thepage \hfil \leftmark \strut}\hrule}}}
40: \newcommand{\heads}[2]{\markboth{\protect\small\it #1}{\protect\small\it #2}}
41: \newcommand{\Acknow}[1]{\subsection*{Acknowledgement} #1}
42:
43: %% TITLE BLOCK
44:
45: \newcommand{\Arthead}[3]{ \setcounter{page}{#2}\thispagestyle{empty}\noi
46: \unitlength=1pt \begin{picture}(500,40)
47: \put(0,58){\shortstack[l]{\small\it Gravitation \& Cosmology,
48: \small\rm Vol. 5 (1999), No. #1, pp. #2--#3\\
49: \footnotesize \copyright \ 1999 \ Russian Gravitational Society} }
50: \end{picture} }
51: \newcommand{\Title}[1]{\noi {\Large #1} \\}
52: \newcommand{\Author}[2]{\noi{\large\bf #1}\\[2ex]\noindent{\it #2}\\}
53: \newcommand{\Authors}[4]{\noi
54: {\large\bf #1\dag\ #2\ddag}\medskip\begin{description}
55: \item[\dag]{\it #3} \item[\ddag]{\it #4}\end{description}}
56: \newcommand{\Rec}[1]{\noi {\it Received #1} \\}
57: \newcommand{\Recfin}[1]{\noi {\it Received in final form #1} \\}
58: \newcommand{\Abstract}[1]{\vskip 2mm \begin{center}
59: \parbox{16.4cm}{\small\noi #1} \end{center}\medskip}
60: \newcommand{\RAbstract}[3]{ {\bf\noi #1}\\ {\bf\noi #2}
61: \begin{center}\parbox{16.4cm}{\small\noi #3} \end{center}\bigskip}
62: \newcommand{\PACS}[1]{\begin{center}{\small PACS: #1}\end{center}}
63: \newcommand{\foom}[1]{\protect\footnotemark[#1]}
64: \newcommand{\foox}[2]{\footnotetext[#1]{#2}
65: \addtocounter{footnote}{#1} }
66: \newcommand{\email}[2]{\footnotetext [#1]{e-mail: #2}
67: \addtocounter{footnote}{#1} }
68: \def\vlad{Talk presented at the 10th Russian
69: Gravitational Conference, Vladimir, 20-27 June 1999.}
70: \def\vladpl{Plenary talk presented at the 10th Russian
71: Gravitational Conference, Vladimir, 20-27 June 1999.}
72:
73: %% TEXT, SPACES AND FIGURES
74:
75: \newcommand{\Theorem}[2]{\medskip\noi {\bf #1. \ }{\sl #2}\medskip}
76: \newcommand{\Ref}[1]{Ref.\,\cite{#1}}
77: \newcommand{\sect}[1]{Sec.\,#1}
78: \def\nq{\hspace*{-1em}}
79: \def\nqq{\hspace*{-2em}}
80: \def\nhq{\hspace*{-0.5em}}
81: \def\nhh{\hspace*{-0.3em}}
82: \def\cm{\hspace*{1cm}}
83: \def\inch{\hspace*{1in}}
84: \def\wide{\mbox{$\dst\vphantom{\int}$}}
85: \def\Wide{\mbox{$\dst\vphantom{\intl_a^a}$}}
86: \def\ten#1{\mbox{$\,\cdot\, 10^{#1}$}}
87: \def\deg{\mbox{${}^\circ$}\ } %% degree
88: \def\degC{\mbox{${}^\circ$C}\ } %% Celsius degree
89: \newcommand{\Figure}[2]{\begin{figure}
90: \framebox[83mm]{\rule{0cm}{#1}}
91: \caption{\protect\small #2}\medskip\hrule\end{figure}}
92:
93: %% EQUATIONS %%
94:
95: %% aligning, numbering
96:
97: \def\al{&\nhq}
98: \def\lal{&&\nqq {}}
99: \def\eq{Eq.\,}
100: \def\eqs{Eqs.\,}
101: \def\beq{\begin{equation}}
102: \def\eeq{\end{equation}}
103: \def\bear{\begin{eqnarray}}
104: \def\bearr{\begin{eqnarray} \lal}
105: \def\ear{\end{eqnarray}}
106: \def\earn{\nonumber \end{eqnarray}}
107: \def\nn{\nonumber\\ {}}
108: \def\nnv{\nonumber\\[5pt] {}}
109: \def\nnn{\nonumber\\ \lal }
110: \def\nnnv{\nonumber\\[5pt] \lal }
111: \def\yy{\\[5pt] {}}
112: \def\yyy{\\[5pt] \lal }
113: \def\eql{\al =\al}
114: \def\sequ#1{\setcounter{equation}{#1}}
115:
116: %% fractions
117:
118: \def\dst{\displaystyle}
119: \def\tst{\textstyle}
120: \def\fracd#1#2{{\dst\frac{#1}{#2}}}
121: \def\fract#1#2{{\tst\frac{#1}{#2}}}
122: \def\Half{{\fracd{1}{2}}}
123: \def\half{{\fract{1}{2}}}
124:
125: %% other
126:
127: \def\eqdef{\stackrel{\rm def}=}
128: \def\e{{\,\rm e}}
129: \def\d{\partial}
130: \def\re{\mathop{\rm Re}\nolimits}
131: \def\im{\mathop{\rm Im}\nolimits}
132: \def\arg{\mathop{\rm arg}\nolimits}
133: \def\tr{\mathop{\rm tr}\nolimits}
134: \def\sign{\mathop{\rm sign}\nolimits}
135: \def\diag{\mathop{\rm diag}\nolimits}
136: \def\dim{\mathop{\rm dim}\nolimits}
137: \def\const{{\rm const}}
138:
139: \def\then{\ \Rightarrow\ }
140: \def\iff{\ \Longleftrightarrow\ }
141: \newcommand{\toas}{\mathop {\ \longrightarrow\ }\limits }
142: \newcommand{\aver}[1]{\langle \, #1 \, \rangle \mathstrut}
143: \def\DAL{\mathop{\raisebox{3.5pt}{\large\fbox{}}}\nolimits}
144: \newcommand{\vars}[1]{\left\{\begin{array}{ll}#1\end{array}\right.}
145: \newcommand{\lims}[1]{\mathop{#1}\limits}
146: \newcommand{\limr}[2]{\raisebox{#1}{${\lims{#2}}$}}
147: \def\suml{\sum\limits}
148: \def\intl{\int\limits}
149:
150: \makeatother
151:
152:
153: \newcommand{\Picture}[4]{
154: \begin{figure} \centering \unitlength=1mm
155: \begin{picture}(82.5,#2)
156: \put(0,0){\line(0,1){#2}} %frame
157: \put(0,0){\line(1,0){82.5}}
158: \put(82.5,0){\line(0,1){#2}}
159: \put(0,#2){\line(1,0){82.5}}
160: \put(#1,#2){#3} \end{picture}
161: \caption{\protect\small #4} \medskip \hrule \end{figure} }
162:
163: \newcommand{\WPicture}[4]{
164: \begin{figure*} \centering \unitlength=1mm
165: \begin{picture}(174,#2)
166: \put(0,0){\line(0,1){#2}} %frame
167: \put(0,0){\line(1,0){174}}
168: \put(174,0){\line(0,1){#2}}
169: \put(0,#2){\line(1,0){174}}
170: \put(#1,#2){#3} \end{picture}
171: \caption{\protect\small #4} \medskip \hrule \end{figure*} }
172:
173: \renewcommand\floatpagefraction{.9} %
174: \renewcommand\topfraction{.9} %
175: \renewcommand\bottomfraction{.9} % To place tall
176: \renewcommand\textfraction{.1} % figures
177: \renewcommand\dblfloatpagefraction{.9} % in the text
178: \renewcommand\dbltopfraction{.9} %
179:
180: \def\Gun{\mbox{${\rm N}\cdot{\rm m}^2/{\rm kg}^2$}}
181: \def\flun{\mbox{${\rm cm}^{-2}{\rm s}^{-1}$}\ } %% flux unit
182: \def\qg{\mbox{${\rm CGSE}_q$}}
183:
184: \def\ME{M_{\oplus}}
185: \def\RE{R_{\oplus}}
186: \def\Horb{H_{\rm orb}}
187:
188: \def\rt{\vec r{}^{\rm \ th}}
189: \def\xt{x{}^{\rm th}}
190: \def\yt{y{}^{\rm th}}
191:
192: \def\gsim{\mathrel{
193: \raisebox{3pt}{$\mathop{>}\limits_{\displaystyle \sim}$}
194: }}
195: \def\lsim{\mathrel{
196: \raisebox{3pt}{$\mathop{<}\limits_{\displaystyle \sim}$}
197: }}
198: \def\z{\mbox{$\phantom{0}$}}
199: \def\zz{\mbox{$\phantom{.0}$}}
200: \def\gapup{\vphantom{\raisebox{1.5mm}{0}}}
201: \def\wider{\vphantom{\raisebox{1.3mm}{0}\raisebox{-1.3mm}{0}}}
202:
203: \renewcommand{\thetable}{\arabic{table}}
204:
205: \heads
206: {A.D. Alexeev, K.A. Bronnikov, N.I. Kolosnitsyn,
207: M.Yu. Konstantinov, V.N. Melnikov and A.J. Sanders}
208: {SEE Project: Current Status and New Estimates}
209:
210: %\bls{0.99}
211:
212: \begin{document}
213: \twocolumn[
214: \Arthead{1 (17)}{67}{78}
215:
216: \Title{SEE PROJECT FOR TESTING GRAVITY IN SPACE:\yy
217: CURRENT STATUS AND NEW ESTIMATES}
218:
219: \Authors{A.D. Alexeev, K.A. Bronnikov, N.I. Kolosnitsyn,
220: M.Yu. Konstantinov, \\ V.N. Melnikov\foom 1}
221: {and A.J. Sanders\foom 2}
222: {RGS, 3-1 M. Ulyanovoy St., Moscow 117313, Russia}
223: {Dept. of Physics and Astronomy, University of Tennessee,
224: Knoxville, TN 37996-1200, USA}
225:
226: \Rec{20 December 1998}
227:
228: \Abstract
229: {We describe some new estimates concerning the recently proposed SEE
230: (Satellite Energy Exchange) experiment for measuring the gravitational
231: interaction parameters in space. The experiment entails precision
232: tracking of the relative motion of two test bodies (a heavy ``Shepherd",
233: and a light ``Particle") on board a drag-free capsule. The new estimates
234: include (i) the sensitivity of Particle trajectories and $G$ measurement
235: to the Shepherd quadrupole moment uncertainties; (ii) the measurement
236: errors of $G$ and the strength of a putative Yukawa-type force whose
237: range parameter $\lambda$ may be either of the order of a few metres or
238: close to the Earth radius; (iii) a possible effect of the Van Allen
239: radiation belts on the SEE experiment due to test body electric
240: charging. }
241:
242: ] %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
243: \email 1 {melnikov@rgs.phys.msu.su}
244: \email 2 {asanders@utkux.utcc.utk.edu}
245:
246:
247: \section {Introduction}
248:
249: The SEE (Satellite Energy Exchange) concept of a space-based
250: gravitational experiment was suggested in the early 90s \cite{SD92}
251: and was aimed at precisely measuring the gravitational interaction
252: parameters: the gravitational constant $G$, possible
253: violations of the equivalence principle measured by the
254: E\"otv\"os parameter $\eta$, time variations of $G$, and
255: hypothetical non-Newtonian gravitational forces
256: (parametrized by the Yukawa strength $\alpha$ and range
257: $\lambda$). Such tests are intended to fill gaps left by current
258: methods of ground-based experimentation and observation of astronomical
259: phenomena. The significance of new measurements is quite evident
260: since nearly all modified theories of gravity and unified
261: theories predict some violations of the Equivalence Principle
262: (EP), either by deviations from the Newtonian law
263: (inverse-square-law, ISL) or by composition-dependent (CD) gravity
264: accelerations, due to the appearance of new possible massive particles
265: (partners); time variations of $G$ are also generally predicted
266: \cite{dSMP, Mel94}.
267:
268: The idea of the SEE method is to study the relative motion of two
269: bodies on board a drag-free Earth satellite using horseshoe-type
270: trajectories, previously known in planetary satellite astronomy:
271: if the lighter body (the "Particle") is moving along a lower
272: orbit than the heavier one (the "Shepherd") and approaching
273: from behind, then the Particle almost overtakes the Shepherd,
274: but it gains energy due to their
275: gravitational interaction, passes therefore to a higher orbit and
276: begins to lag behind. The interaction phase can be studied within a
277: drag-free capsule (a cylinder up to 20 m long, about 1 m in diameter)
278: where the Particle can loiter as long as $10^5$ seconds.
279: It was claimed that the SEE method exceeded in accuracy all other
280: suggestions, at least with respect to $G$ and $\alpha$ for $\lambda$ of
281: the order of metres. Some design features were considered, making it
282: possible to reduce various sources of error to a negligible level.
283: It was concluded, in particular, that the most favourable orbits are
284: the sun-synchronous, continuous sunlight orbits situated at altitudes
285: between 1390 and 3330 km.
286:
287: Since the origination of the SEE concept, the development
288: has focused on critical analyses and critical hardware requirements.
289: All indications from this work are that the SEE concept is feasible and
290: practicable \cite{iztech}. A ``blue ribbon" Theory Advisory Group,
291: formed two years ago to critique Project-SEE activities and goals, has
292: concluded that they are sound.
293:
294: This paper presents some new evaluations concerning the opportunities of
295: the SEE concept and its yet-unresolved difficulties. In \sect 2, for
296: comparison, we briefly outline the current status of terrestrial and
297: astronomical determination of the gravitational interaction parameters
298: to be measured by the SEE method. In \sect 3, on the basis of computer
299: simulations of Particle trajectories, we estimate the requirements for
300: the Shepherd quadrupole moment uncertainty. \sect 4 shows the results of
301: simulations of the measurement procedure itself, which enables us to
302: estimate the possible measurement accuracy with respect to $G$ and
303: $\alpha$ for $\lambda$ of the order of either metres or the Earth's
304: radius. \sect 5 discusses a spurious effect of test body electric
305: charging when the satellite orbit passes through the Van Allen radiation
306: belts, rich in high-energy protons. \sect 6 is a conclusion.
307:
308: In what follows, the term ``orbit" applies to satellite (or
309: Shepherd) motion around the Earth, while the words ``trajectory" or
310: ``path" apply to Particle motion with respect to the Shepherd inside the
311: capsule.
312:
313: \section{State of the art: a brief survey}
314:
315: Since gravitational forces are so very small, precision-measurement
316: techniques have been at the core of terrestrial gravity research for two
317: centuries. However, evidence is increasingly accumulating which
318: indicates that terrestrial methods have plateaued in accuracy and are
319: unlikely to achieve significant accuracy gains in the future
320: \cite{Gi97}. For example, the uncertainty in the gravitational constant
321: $G$ had been accepted as 128 ppm for nearly two decades, and the actual
322: uncertainty in $G$ --- as indicated by the scatter of results among
323: recent experiments which claim high accuracy --- is roughly the same
324: (about 140 ppm). We discuss below the situation with respect to several
325: key measurements.
326:
327: \subsection{Terrestrial determinations of $G$}
328:
329: The Luther \& Towler determination of $G$ in 1982 \cite{LuTh82},
330: with the result $(6.6726\pm 0.0005) \ten{-11}$ \Gun\ and other,
331: less precise experiments gave rise to the current official CODATA value
332: of $G$, viz. 6.67259\ten{-11} \Gun\ with an error of 128
333: ppm. Several still other experiments which also claimed high precision
334: were ignored by CODATA because of inadequate documentation of systematic
335: errors.
336:
337: There is considerable evidence that the uncertainty in $G$ has
338: plateaued at about 100 ppm. At a recent (November 23--24, 1998)
339: conference in
340: London, several new (1998) determinations of $G$ were reported. The
341: obtained values for $G$ (in units of $10^{11}$ \Gun) and their estimated
342: error $\delta G/G$ in ppm are as follows:
343: \begin{center}
344: \begin{tabular}{llr}
345: Fitzgerald and Armstrong & 6.6742 & 90
346: \\
347: \cm (New Zealand) \cite{FiArm} & 6.6746 & 134
348: \yy
349: Nolting et al. (Zurich) \cite{Nolting} & 6.6749 & 210
350: \yy
351: Meyer et al. (Wuppethal) \cite{Meyer} & 6.6735 & 240
352: \yy
353: Karagioz et al. (Moscow) \cite{Karag} & 6.6729 & 75
354: \yy
355: CODATA (1986) & 6.67259 & 128
356: \end{tabular}
357: \end{center}
358:
359: Obviously, most of the stated errors are of the order 100 ppm. Moreover,
360: the scatter (1-sigma) about the mean is about 140 ppm. Some of the
361: investigators still hope for accuracy of 10 ppm. It remains to be seen
362: whether they will be able to report error estimates of this size or,
363: more importantly, whether their respective values for $G$ will actually
364: agree within 10 ppm.
365:
366: We note that this analysis is perhaps unduly optimistic since it
367: excludes one extremely bad outlier: the very careful and well documented
368: experiment by the Physikalisch-Technische Bundesanstalt in the late
369: 1980s and early 1990s, which obtained a series of values for $G$ that
370: were consistenly above the results of most experimenters by about 6000
371: ppm (0.6\,\% !), while claiming an error of about 100 ppm \cite{PTB}. No
372: explanation for such a large discrepancy has been found.
373:
374: It might seem that the problems of terrestrial apparatus must inexorably
375: yield to new technologies --- that the promise of ever increasing
376: sensitivities would also lead to ever improving accuracy. However, this
377: may not be true, since it is various systematic errors which limit the
378: ultimate attainable accuracy in terrestrial experiments \cite{Gi97}.
379:
380:
381: \subsection{Terrestrial tests of the equivalence principle (EP) and search
382: for Yukawa forces}
383:
384:
385: The EP may be tested by searching for either violations of the
386: inverse-square law (ISL) or composition-dependent (CD) effects in
387: gravitational free fall.
388:
389: In the watershed year of 1986, Fischbach startled the physics
390: community by showing that E\"otv\"os's famous turn-of-the-century
391: experiment is much less decisive as a null result than was generally
392: believed \cite{Fis86}. Prior to this time, experiments by
393: Dicke \cite{Roll64} and Braginsky \cite{Brag71} had demonstrated the
394: universality of free fall (UFF) to very high accuracy with respect to
395: several metals falling in the gravitational field of the Sun
396: (the E\"otv\"os parameter $\eta$ was ultimately found to be smaller than
397: $10^{-12}$). The interpretation of these results at the time was that
398: they validated UFF.
399:
400: It was implicit that any violation would have infinite range, like
401: gravity \cite{Adel94}. During the 1970s and early 1980s there was also a
402: flurry of activity concerning possible ISL violations, which eventually
403: led to null results at the levels of precision then available (Fujii,
404: \cite{Fuj71-72}, Long \cite{Long76-84}).
405:
406: Since 1986 it has become customary to parametrize possible apparent
407: EP violations as if due to a Yukawa particle with a Compton wavelength
408: $\lambda$. This approach unites both ISL and CD effects very naturally,
409: while the parameter values in the Yukawa potential suggest which
410: experimental conditions are required to detect the new interaction.
411:
412: Following Fischbach's conjecture, ISL and CD tests were
413: undertaken by many investigators. Although a number of anomalies
414: were initially reported, nearly all of these were eventually explained in
415: terms of overlooked systematic errors or extreme sensitivity to models,
416: while most investigators obtained null results. By far the tightest
417: bounds are those obtained by Adelberger and his ``Eot-Wash" group at
418: the University of Washington \cite{Adel+}.
419: This group expects a further improvement of at least an order of
420: magnitude \cite{Adel97}.
421: A positive result for a deviation from the Newtonian law (ISL) was
422: obtained (and interpreted in terms of a Yukawa-type potential) in the
423: range of 20 to 500 m by Achilli and colleagues \cite{Ach}; this needs to
424: be verified in other independent experiments.
425:
426: For reviews of terrestrial searches for non-Newtonian gravity, see
427: \cite{Adel94, FiG, Franklin}. The opportunities of the SEE concept in
428: this respect are discussed in Refs.\,\cite{SD92, iztech} and in the
429: present paper.
430:
431: The UFF is still in the scope of the current experimental
432: projects, and the SEE concept suggests here a progress of 3
433: to 4 orders of magnitude as compared with \Ref{Brag71}. Only one
434: project, STEP (Satellite Test of the EP) promises a greater
435: progress but meets some significant problems of its own \cite{STEP},
436: connected, in particular, with the radiation belts.
437:
438: \section
439: {Simulations of Particle trajectories and the Shepherd quadrupole moment}
440:
441: In the previous studies of the SEE project it was assumed that the capsule
442: was about 20 m long and the initial Shepherd-Particle separation $x_0$
443: along the capsule axis was as
444: great as 18 m; some estimations were also made for $5$ m $\leq x_0\leq
445: 10$ m. The Shepherd mass was taken to be $M=500$ kg and the Particle mass
446: $m=0.1$ kg. The present study retains these values.
447:
448: In what follows we describe some characteristic features of Particle
449: trajectories with a goal to determine
450: their sensitivity to the uncertainty of the Shepherd quadrupole
451: moment $J_2$ for $x_0\geq 5$ m.
452: As in our previous studies, the capsule diameter is supposed to be 1 m.
453:
454: The reason for considering the quadrupole moment uncertainty is
455: technological by origin. Namely, it is hard to produce a spherically
456: symmetric Shepherd to a required accuracy and, instead, it has been
457: suggested \cite{SD92} to use a Cook-Marussi stack of cylinders with $J_2
458: >0$, which may be manufactured more easily. A slow rotation of the Shepherd
459: with $J_2>0$ will stabilize its position and orientation.
460:
461: The value of $J_2$ can be provided with some uncertainty $\delta J_2$.
462: To avoid the inclusion of $\delta J_2$ in the set of parameters to be
463: determined in the experiment, it is useful to know which values of $\delta
464: J_2$ will be negligible, since the growth of the number of parameters
465: leads to serious problems in data processing.
466:
467:
468: \subsection{Equations of motion and the initial data}
469:
470:
471: Assuming that the relative motion of the test bodies inside the capsule
472: occurs in the satellite orbital plane,
473: the reduced Lagrangian of the Particle motion reads
474: \bearr
475: \label{lagr}
476: L=\frac M2(\dot R^2+R^2\dot \varphi ^2) \nnn
477: \cm +\frac m2\left[ \dot r^2+r^2(\dot \varphi +\dot \psi )^2\right]
478: +G\frac{\ME m}r \nnn
479: \nq\ +G\frac{Mm}s\left\{ 1+J_2\left(\frac{r_s}s\right)^2\!
480: P_2(\cos\theta )\right\}
481: \left( 1+\alpha e^{-s/\lambda }\right)
482: \ear
483: where $( R,\varphi)$ are the Earth-centred polar coordinates of the
484: Shepherd in the orbital plane;
485: $r=\sqrt{(R{+}y)^2{+}x^2}$ and $\psi$
486: are the Earth-centred polar coordinates of the Particle;
487: $x$ and $y$ are the Shepherd-centred Particle coordinates, where $x$ is the
488: ``horizontal" one, i.e., along the orbit and simultaneously along the
489: capsule and $y$ is the ``vertical" one, along the Earth-Shepherd radius
490: vector;
491: $s=\sqrt{x^2+y^2}$ is the Particle-Shepherd separation;
492: $\ME $, $M$ and $m$ are the Earth, Shepherd and Particle masses,
493: respectively;
494: $J_2$ is the quadrupole moment of the Shepherd, $r_s$ is its
495: radius and $P_2$ is the Legendre polynomial
496: \[
497: P_2(\cos \theta )=\frac{3\cos ^2\theta -1}2,
498: \]
499: where $\theta$ is the angle between the line connecting the centres of
500: the test bodies and the Shepherd equatorial plane. It is easy to see
501: that if the Shepherd symmetry axis is in its orbital plane, then
502: $\theta =\theta _0=-\arctan (y/x) +\varphi $. If the symmetry
503: axis of Shepherd is orthogonal to its orbital plane, then $\theta =0$. In
504: general, if $\chi$ is the angle between the Shepherd symmetry axis
505: and its orbital plane, then $\theta =\theta _0\cos \chi $. Hence the
506: influence of $J_2$ on the Particle motion is
507: minimum if the Shepherd symmetry axis lies in its orbital plane
508: and is maximum if they are mutually orthogonal.
509:
510: For simplicity (and taking into account the corresponding estimate)
511: we neglect the influence of the Particle on the Shepherd, so the
512: Shepherd trajectory is considered to be given. Then, varying
513: the above Lagrangian with respect to $x$ and $y$, taking into account that
514: $M\gg m$ and $R\gg s$, we arrive at the following equations of Particle
515: motion with respect to the Shepherd:
516: \def\M{\overline{M}}
517: \bearr
518: \frac{d^2x}{dt^2}=2\dot y\dot \varphi +x\left\{ \dot \varphi ^2-\frac{G\ME}{
519: r^3}\right\} -\frac{2\dot R\dot \varphi y}R
520: \nnn
521: -\frac{G\M}{s^3}x\left\{
522: 1+J_2\left( \frac{r_s}s\right) ^2\! P_2(\cos \theta )\right\}
523: \nnn
524: -\alpha x\frac{G\M}{s^2}\left\{ 1+J_2\left( \frac{r_s}s\right) ^2\! P_2(\cos
525: \theta )\right\}\! \left(\frac 1s+\frac 1\lambda \right) \e^{-s/\lambda}
526: \nnn \label{eqnx}
527: -\frac{G\M r_0^2}{2s^5}J_2\left( 1+\alpha e^{-s/\lambda }\right) \times
528: \nnn \cm
529: \times \left[ x(1+3\cos 2\theta )+3y\sin 2\theta \cos \chi \right]
530: \yyy
531: \frac{d^2y}{dt^2}=-2\dot x\dot \varphi +(R+y)\left\{ \dot \varphi ^2-\frac{
532: G\ME }{r^3}\right\} +\frac{2\dot R\dot \varphi x}R
533: \nnn
534: -\frac{G\M}{s^3}y\left\{
535: 1+J_2\left( \frac{r_s}s\right) ^2\! P_2(\cos \theta )\right\}
536: \nnn
537: -\alpha y\frac{G\M}{s^2}\left( \frac 1s+\frac 1\lambda \right) \!\left\{
538: 1+J_2\left( \frac{r_s}s\right) ^2\! P_2(\cos \theta )\right\} e^{-s/\lambda}
539: \nnn \label{eqny}
540: +\frac{G\M r_0^2}{s^5}J_2\left( 1+\alpha e^{-s/\lambda }\right) \times
541: \nnn \cm
542: \times \left[ 3x\sin 2\theta \cos \chi +y(1-3\cos \theta )\right]
543: \ear
544: where $\M = M+m$.
545:
546: Two kinds of initial conditions for \eqs (\ref{eqnx}) and (\ref{eqny})
547: were used during the simulations. First, we used the so-called ``standard''
548: initial conditions, taking
549: the Particle velocity components $\dot x(0)$ and $\dot y(0)$ corresponding
550: to its unperturbed (i.e., without the $M-m$ interaction) orbital motion
551: distinguished from the Shepherd's orbit only by its radius (for circular
552: orbits) or major semiaxis (for elliptic orbits). Assuming that the Particle
553: motion begins right at the moment when the Shepherd passes its perigee,
554: these conditions have the form
555: \bearr
556: \label{initcond}
557: x(0)=x_0, \cm\cm y(0)=y_0,\nnn
558: \dot x(0)=\frac{\omega
559: e'y_0}{2(1-e)^2},\cm \dot y(0)=-\frac{\omega ex_0}{e'(1-e)}
560: \ear
561: where $\omega ^2=G\ME /R_0^3$, $R_0$ is the Shepherd orbital
562: radius (at the perigee), $e$ is the orbital eccentricity and $e'=\sqrt{1-e^2}$.
563:
564: For clearness, the relations (\ref{initcond}) are written in the linear
565: approximation in the variables $x$ and $y$. Higher-order approximations were
566: used in the simulation process as well.
567:
568: The second kind of initial conditions correspond to small variations of
569: initial velocities with respect to their ``standard'' values.
570:
571: The set of equations (\ref{eqnx})--(\ref{eqny}) was solved numerically using
572: the software developed previously \cite{iztech} to analyze the SEE project.
573:
574: On the basis of numerical solution of \eqs (\ref{eqnx}) and (\ref{eqny}), we
575: considered two types of Particle trajectories, corresponding to different
576: choices of the initial data: (i) approximately U-shaped ones and (ii)
577: cycloidal ones, containing loops (see more details on the trajectories in
578: \cite{SD92,iztech}), for orbital altitudes $\Horb=500$, 1500 and 3000 km.
579: The uncertainty $\delta J_2$ ranged in the interval $10^{-3}\div 10^{-5}$;
580: and initial Particle position changed in the range $6$ m $\leq x_0\leq 18$
581: m, $-25$ cm $\leq y_0\leq -5$ cm.
582:
583: For $\Horb=500$ km, all U-shaped paths contained a sinusoidal component
584: (with the orbital frequency), starting at $x_0\leq 8$ m, while
585: for $ \Horb=1500$ and $3000$ km it was present in paths
586: starting at $ x_0\leq 10$ m. In other families of U-shaped trajectories
587: a sinusoidal component was present only in the case $|y_0|\geq 20$ cm.
588:
589:
590: \subsection{Restrictions on the Shepherd quadrupole moment uncertainty}
591:
592:
593: Small Shepherd quadrupole moment uncertainties $\delta J_2$ create
594: small displacements $\delta\vec r$ of a Particle trajectory
595: with respect to unperturbed one, $\vec r_0(t)$:
596: \[
597: \delta \vec{r}= \vec r_j(t)-\vec r_0(t)
598: \]
599: where $\vec r_j$ is the perturbed path.
600: Instead of the full displacement $\delta \vec r$, a
601: displacement $\delta x$ along the $x$ axis may be considered
602: since, by numerical simulations, displacements along the $y$ axis
603: are an order of magnitude smaller than $\delta x$.
604:
605: Numerical simulations show that in the whole range of the above initial
606: conditions the displacement $\delta x$ is (as it should naturally be) a
607: linear function of $\delta J_2$; for $\delta J_2=10^{-4}$.
608: For the case when the Shepherd symmetry axis is located in the orbital
609: plane, the maximum values of $\delta x$ for U-shaped trajectories are given
610: in Table 1.
611:
612: \begin{table}
613: \noi
614: {\bf Table 1.}\
615: Displacements of U-shaped trajectories under $\delta J_2=10^{-4}$ for
616: the Shepherd symmetry axis in its orbital plane.
617: The second line shows $x_0$.
618:
619: \begin{center}
620: \begin{tabular}{|r|r|r|r|r|r|}
621: \hline \wide
622: $y_0$ & \multicolumn{5}{|c|}{$\wide \delta x_{\max} \times 10^{7}$ (m)}\\
623: \cline{2-6} \wide
624: (cm) & 18 m & 10 m & 8 m & 6 m & 4 m \\
625: \hline \gapup
626: -25 & 13.3\z & 7.52 & 7.98 & 8.57 & 14.90 \\
627: -20 & 5.51 & 4.65 & 4.83 & 5.51 & 5.84 \\
628: -15 & 1.98 & 2.87 & 3.03 & 3.3\z & 2.18 \\
629: -10 & 0.56 & 1.59 & 1.7\z & 1.88 & 0.66 \\
630: - 5 & 0.11 & 0.59 & 0.62 & 0.69 & 0.14 \\
631: \hline
632: \end{tabular}
633: \end{center}
634: \end{table}
635:
636: One can conclude that, if the distance measurement error is $10^{-6}$ m,
637: for most of the trajectories the uncertainty $\delta J_2 = 10^{-4}$ is
638: admissible.
639:
640: The increase of $\delta x$ for small values of $|y_0|$ is explained
641: by a large displacement of the turning point towards the Shepherd.
642:
643: When the Shepherd symmetry axis is orthogonal to the orbital plane, these
644: estimations change as shown in Table 2.
645:
646: \begin{table}
647: \noi {\bf Table 2.}
648: Displacements of U-shaped trajectories under $\delta J_2=10^{-4}$ for
649: the Shepherd symmetry axis orthogonal to its orbital plane.
650: The second line shows $x_0$.
651:
652: \begin{center}
653: \begin{tabular}{|r|r|r|r|r|r|}
654: \hline \wide
655: $y_0$ & \multicolumn{5}{|c|}{$\wide \delta x_{\max} \times 10^{7}$ (m)}\\
656: \cline{2-6} \wide
657: (cm) & 18 m & 10 m & 8 m & 6 m & 4 m \\
658: \hline \gapup
659: -25 & 54.2\z & 62.4\z & 68.9\z & 81\z\z\, & 57\z\z\,\\
660: -20 & 21.9\z & 25.8\z & 28.1\z & 32.2\z & 23.3\z \\
661: -15 & 7.94 & 10.2\z & 11.5\z & 14\z\z\, & 8.7\z \\
662: -10 & 2.25 & 3.42 & 4.15 & 5.45 & 2.65 \\
663: - 5 & 0.45 & 0.83 & 1.05 & 1.47 & 0.58 \\
664: \hline
665: \end{tabular}
666: \end{center}
667: \end{table}
668:
669: For cycloidal trajectories these values are approximately an order
670: of magnitude smaller than those for the U-shaped ones.
671: % , see table 3,
672: % where values of $\delta x$ for cycloidal trajectories are given
673: % for only two values of $x_0$ but for both
674: % orientations of the Shepherd symmetry axis.
675: %
676: % \begin{table}
677: % {\bf Table 3.}
678: % Displacements of cycloidal trajectories for $\delta J_2=10^{-4}$.
679: %
680: % \begin{center}
681: % \begin{tabular}{|r|r|r|r|r|}
682: % \hline
683: % & \multicolumn{4}{|c|}{$\wide \delta x_{\max} \times 10^{8}$ (m)}\\
684: % \cline{2-5}
685: % $y_0$ \ \ & \multicolumn{2}{|c|}{$x_0 = 18 $m}
686: % & \multicolumn{2}{|c|}{$x_0=6$ m} \\
687: % \cline{2-5}
688: % (cm) & $\chi =0$ & $\chi =\pi/2$ & $\chi =0$ & $\chi =\pi/2$ \\
689: % \hline
690: % -25 & 8.05 & 32\z\zz & 16.4\z & 71\z\zz \\
691: % -20 & 0.95 & 3.83 & 2.93 & 12\z\zz \\
692: % -15 & 0.26 & 1.07 & 0.04 & 0.12 \\
693: % \hline
694: % \end{tabular}
695: % \end{center}
696: % \end{table}
697:
698: It was also found that $\delta x$ decreases with increasing
699: orbital altitude $\Horb$. Table 3 shows, as an example, the displacements of
700: U-shaped trajectories with $x_0=18$ m for $\Horb=500$, 1500 and 3000 km
701: and the Shepherd symmetry axis located in its orbital plane.
702:
703: \begin{table}
704:
705: {\bf Table 3.}
706: Displacements of U-shaped trajectories with $x_0=18$ m and different
707: orbital altitudes for $\delta J_2=10^{-4}$. The second line shows the
708: values of $\Horb$.
709:
710: \begin{center}
711: \begin{tabular}{|r|r|r|r|}
712: \hline
713: $y_0$ & \multicolumn{3}{|c|}
714: { \wide $\delta x_{\max} \times 10^{7}$ (m)} \\
715: \cline{2-4}
716: (cm) & 500 km & 1500 km & 3000 km \\
717: \hline
718: -25 & 42.6\z & 13.3\z & 4.84 \\
719: -20 & 11.9\z & 5.51 & 2.18 \\
720: -15 & 4.02 & 1.98 & 0.86 \\
721: -10 & 1.03 & 0.56 & 0.28 \\
722: -5 & 0.17 & 0.11 & 0.07 \\
723: \hline
724: \end{tabular}
725: \end{center}
726: \end{table}
727:
728: The above results show that the Shepherd quad\-ru\-pole moment uncertainty
729: $\delta J_2$ may be neglected in the SEE experiment with circular Shepherd
730: orbits at $\Horb=1500$ or $3000$ km and U-shaped Particle trajectories if
731: $\delta J_2\leq 10^{-5}$ and the position measurement error $\delta l$ is
732: $10^{-6}$ cm, or $\delta J_2\leq 10^{-7}$ for $\delta l = 10^{-8}$ cm. For
733: cycloidal Particle trajectories or elliptic Shepherd orbits these estimates
734: become $\delta J_2\leq 10^{-4}$ and $\delta J_2\leq 10^{-6}$, respectively.
735: For low orbits, $\Horb=500$ km, the resitrictions on $\delta J_2$ become
736: more stringent: $\delta J_2\leq 10^{-6}$ for $\delta l = 10^{-6}$ cm and
737: $\delta J_2\leq 10^{-8}$ for $\delta l = 10^{-8}$ cm. However, as is evident
738: from the above tables, these requirements may be relaxed by an order of
739: magnitude if one discards some trajectories.
740:
741: The influence of $\delta J_2$ on the accuracy of $G$ measurement may be now
742: estimated as follows. Let some value of $\delta J_2$
743: produce the trajectory displacement $| \delta \vec{r}|
744: \leq \delta l_j$ while the variation $\delta G_0$ of $G$ with the same
745: initial conditions gives the trajectory displacement
746: $|\delta\vec {r}| \leq \delta l_{G}$. Then, keeping in mind the
747: linear dependence of trajectory displacements on $\delta J_2$ and
748: $\delta G$, the accuracy of $G$ measurement under the
749: uncertainty $\delta J_2$ may be estimated as
750: \[
751: \frac{\delta G}G\leq \frac{\delta l_j}{\delta l_{G}}\frac{\delta G_0}G.
752: \]
753:
754: Using this inequality and the results of trajectory simulations,
755: we obtain the following estimates for U-shaped Particle trajectories
756: in circular orbits with $\Horb=1500$ km:
757:
758: \medskip\noi
759: {\bf Table 4. } Estimates of $\delta G/G$ in ppm
760: for $\delta J_2=10^{-4}$, when the symmetry axis of the Shepherd lies in
761: ($\chi=0$) or is ortogonal to ($\chi=\pi/2$) its orbital plane.
762: The second line shows $x_0$.
763:
764: \begin{center}
765: \begin{tabular}{|r|r|r|r|r|}
766: \hline
767: $y_0,$ &
768: \multicolumn{2}{|c|}{$\chi=0$}&
769: \multicolumn{2}{|c|}{$\chi=\pi/2$} \\
770: \cline{2-5}
771: cm & 18 m & 6 m & 18 m & 6 m \\
772: \hline
773: -25 & $0.88$ & $0.7$ & $3.57$ & $6.7\z$ \\
774: \hline
775: -20 & $0.27$ & $0.3$ & $1.05$ & $1.73$ \\
776: \hline
777: -15 & $0.06$ & $0.1$ & $0.24$ & $0.44$ \\
778: \hline
779: \end{tabular}
780: \end{center}
781:
782: One can conclude that the uncertainties $\delta J_2 \lsim 10^{-5}$
783: do not create substantial $G$ errors for most of the trajectories.
784:
785:
786: \section {Simulations of experimental procedures}
787:
788:
789: This section describes the results of computer simulations of the whole
790: measurement procedures aimed at obtaining the sought-after gravitational
791: interaction parameters. These simulations assumed the
792: Shepherd mass $M=500$ kg, a circular orbit with $\Horb=1500$ km under a
793: spherical gravitational potential of the Earth, and a Particle mass of $100$
794: g. Where relevant, it is assumed that both the Shepherd and the Particle are
795: made of tungsten. Their identical compositions are assumed for simplicity
796: since this work is performed only for estimation purposes.
797:
798:
799: \subsection{Equations of motion with Yukawa terms}
800:
801:
802: We will begin with a presentation of the Particle equations of motion in the
803: relevant approximation, including the contributions from hypothetical Yukawa
804: forces, taking into account the finite size of the Yukawa field sources.
805:
806: Let the interaction potential for two elementary masses $m_1$ and $m_2$
807: be described by the potential
808: \beq \label{Yuk1}
809: dV^{\rm Yu} = \frac{G\,dm_1 dm_2}{r} \alpha\e^{-r/\lambda}
810: \eeq
811: where $r$ is the masses' separation, $\alpha$ and $\lambda$ are the
812: strength parameter and the range of the Yukawa forces. Then for two
813: massive bodies with the radii $R_1$ and $R_2$
814: after integration over their volumes we obtain \cite{ZaKol}
815: \beq
816: V^{\rm Yu} = \frac{G\,m_1 m_2 \beta_1 \beta_2}{r} \label{Yuk2}
817: \alpha\e^{-r/\lambda}
818: \eeq
819: where
820: \beq \label{Yuk3}
821: \beta_i = 3 \biggl(\frac{\lambda}{R_i}\biggr)^3
822: \biggl[\frac{R_i}{\lambda}\cosh \frac{R_i}{\lambda}
823: - \sinh \frac{R_i}{\lambda} \biggr].
824: \eeq
825: When $R_i/\lambda \ll 1$, we have $\beta_i \approx 1$.
826: This may be the case when we consider the interaction between the
827: Shepherd and the Particle at a distance of the order of a few metres.
828: The radii of the Shepherd and the Particle are small: $R_1 \approx
829: 18$ cm for the Shepherd and $R_2 \approx 1.1$ cm for the Particle.
830: If the range $\lambda$ is of the order of the Earth radius, $\lambda
831: \approx \RE$, we have $\beta_{\oplus}=1.10$ and $\beta_{1,2}=1$ where
832: the indices 1 and 2 label the Shepherd and the Particle, respectively.
833:
834: The equations of motion are obtained under
835: the following assumptions. There are two Yukawa interactions
836: with the parameters $\lambda_0$ and $\alpha_0$
837: referring to the Earth-Shepherd and Earth-Particle interactions which
838: are the same (due to the assumed identical composition for the Shepherd
839: and the Particle), while $\lambda$ and $\alpha$
840: determine the Shepherd-Particle interaction.
841: The equations of motion in the frame of reference connected with the
842: Shepherd, with the same notations $x$, $y$, $s$ as previously, are
843: \bearr
844: \ddot{x} +2\omega^2 \dot{y}
845: + G(m_1+m_2) \frac{x}{s^3} - 3\omega^2 \frac{xy}{s} \nnn
846: \cm + G(m_1+m_2) \frac{x}{s^3}
847: \alpha \biggl(1+ \frac{s}{\lambda}\biggr)\e^{-s/\lambda}
848: =0; \nnnv
849: \ddot{y}
850: -2 \omega\dot{x} - 3\omega^2 y + G(m_1+m_2)\frac{y}{s^3}
851: + \frac{3\omega^2}{r_{01}} \biggl(y^2 -\frac{x^2}{2}\biggr)\nnn
852: \cm +G(m_1+m_2)\frac{y}{s^3}
853: \alpha \biggl(1+\frac{s}{\lambda}\biggr)\e^{-s/\lambda} \nnn
854: \inch\ \ \ - \omega^2 \beta_0 \alpha_0 \e^{-r_{01}/\lambda_0}y =0
855: \label{Yuk4}
856: \ear
857: where $\omega$ is the orbital frequency:
858: \beq \label{omega}
859: \omega^2 = \frac{G\ME}{r_{01}^3}
860: \biggl[ 1+\beta_0 \alpha_0 \biggl(1+\frac{r_{01}}{\lambda_0}\biggr)
861: \e^{-r_{01}/\lambda_0}\biggr].
862: \eeq
863: We have neglected the terms quadratic in $s/r_{01}$ times $\alpha$ or
864: $\alpha_0$ due to their manifestly small contributions.
865:
866: If we set $\alpha_0=0$ in \eqs (\ref{Yuk4}), we obtain the equations
867: used to describe only the Shepherd-Particle Yukawa interaction.
868: One can notice that Yukawa terms are roughly proportional to the
869: gradients of the corresponding Newtonian accelerations, namely,
870: $Gm_1/s^3$ for the Shepherd-Particle interaction and
871: $G\ME/r_{01}^3 \approx \omega^2$ for (say) the Earth-Shepherd
872: interaction. In our case these quantities are estimated as
873: \bearr
874: \frac{Gm_1}{s^3} \approx 2.7\ten{-10}\ {\rm s}^{-2}
875: \cm {\rm for}\ \ s=5\ {\rm m},\nnn \cm
876: \omega^2 \approx 8.16\ten{-7} \ {\rm s}^{-2}.
877: \ear
878: Thus, given the same strength parameter, the Earth's Yukawa force is
879: three orders of magnitude greater than that between the Shepherd and
880: the Particle, therefore one might expect some significant progress in
881: an ISL test for $\lambda$ of the order of the Earth's radius.
882:
883: The effect of the Earth's Yukawa force is proportional to the
884: displacements of the statellite along the direction of the Earth's
885: radius. Therefore the sensitivity of the SEE method will increase
886: if one uses orbits with eccentricities of the order of 0.01,
887: following Nordtvedt's suggestion \cite{Nor}. (Larger eccentricities
888: would too much disturb the qualitative picture of a SEE encounter.)
889: Tentative estimates show that in this way one can achieve sensitivities
890: to $\alpha \sim 10^{-10}$, and more thourough studies are in progress.
891:
892: \eqs (\ref{Yuk4}) were used to simulate the measurement procedures.
893:
894:
895: \subsection{Simulations of an experiment for measuring $G$}
896:
897:
898: The constant $G$ is determined from the best
899: fitting condition between the ``theoretical''
900: ($\rt (t_i) = \rt_i$) and ``empirical'' ($\vec r_i$) Particle
901: trajectories near the Shepherd. The fitting quality is evaluated by
902: minimizing a functional characterizing a ``distance'' between the
903: trajectories. We have considered the following functionals for such
904: ``distances'':
905: \bearr
906: S = \sum_{i=1}^{N}
907: \biggl [(x_i-\xt_i)^2 + (y_i-\yt_i)^2 \biggr], \label{F1}\\
908: \lal S_x = \sum_{i=1}^{N}
909: (x_i-\xt_i)^2, \qquad
910: S_y = \sum_{i=1}^{N}
911: (y_i-\yt_i)^2, \label{F2} \\
912: \lal S^* = \sum_{i=1}^{N}
913: \biggl [|x_i-\xt_i| + |y_i-\yt_i| \biggr], \label{F3} \\
914: \lal S_x^* = \sum_{i=1}^{N}
915: |x_i-\xt_i|, \qquad
916: S_y^* = \sum_{i=1}^{N}
917: |y_i-\yt_i|. \label{F4}
918: \ear
919: The theoretical trajectory depends on the gravitational constant $G$, on the
920: initial coordinates $x_0, y_0$ and on the initial velocities $v_{x0},
921: v_{y0}$. To estimate $G$, one chooses the value for which a ``distance''
922: functional in the space of the five variables ($G, x_0, y_0, v_{x0}, v_{y0}$)
923: reaches its minimum.
924:
925: % If the coordinate measurement error is distributed
926: % according to the Gaussian law with a dispersion $\sigma^2$, then, at the
927: % minimum point, the ratio $S/\sigma^2$ is distributed according to the
928: % $\chi^2$ law with $2N$ degrees of freedom. This enables one to estimate the
929: % $G$ measurement error $\delta G$ from the data for {\sl one\/}
930: % trajectory. An estimate of $\delta G$ may be also obtained from
931: % a few independent paths.
932:
933: %%\Figure{55mm}
934: \Picture{0}{55}{\special{em:graph pic1.gif}}
935: {Errors $\delta G$ estimated
936: %% from one-path data ($R_{\chi}$),
937: by the gradient descent ($R_{\rm grad}$) and consecutive descent
938: ($R_{\rm s}$) methods}
939:
940: We carried out a computer simulation of the SEE experiment and estimated
941: $\delta G$ for a given coordinate measurement error
942: ($\sigma=1\ten{-6}$ m). As ``empirical'' trajectories, we took computed
943: trajectories, with specified values of the above five variables, where a
944: Gaussian noise was introduced from a random number generator. Independent
945: ``empirical trajectories were created by non-intersecting random number
946: sequences. The functional was minimized using the gradient
947: descent method and the consecutive descent method.
948: The starting value of the ``vertical'' (along the Earth's radius)
949: coordinate, $y_0$, was taken to be 0.25 m, while the horizontal one, $x_0$,
950: varied between 2 and 18 m. Fig\,1 %% Table 5
951: shows the dependence of the errors
952: %% $\delta G/G = R_{\chi}$, obtained from the data of a single trajectory,
953: $\delta G/G = R_{\rm grad}$, obtained by
954: the gradient descent method and $\delta G/G = R_{\rm s}$, obtained by the
955: consecutive descent method. All the errors are estimated by confidence
956: intervals corresponding to a confidence of 0.95. The mean values of these
957: errors are as follows:
958: \bear
959: %% R_{\chi} \eql 6.59\ten{-7},\cm
960: R_{\rm grad} = 4.69\ten{-8},\cm
961: R_{\rm s} \eql 5.24\ten{-8}.
962: \earn
963: Thus the errors estimated by the gradient and consecutive descent methods
964: are close to each other and are about an order of magnitude smaller than the
965: error from one-trajectory data. It has been discovered that the simulation
966: results strongly depend on the random number generator, so that ordinary
967: generators are not perfect.
968:
969: The use of truncated functionals like (2) has shown that a functional
970: incorporating the more informative ``horizontal'' coordinate $x$ leads to
971: estimates close to those obtained from the total functional, whereas the use
972: of $y$ alone substantially decreases the sensitivity. Therefore in
973: practice, to determine $G$, it is sufficient to measure only one of the two
974: coordinates, viz. $x$.
975:
976: Since the ``empirical'' trajectory is built on the basis of a computed one,
977: with a known value of the gravitational constant $G_0$, it appears
978: possible to estimate a possible systematic error inherent in the data
979: processing method. The latter has turned out to be in most cases much
980: smaller than the random error. This result shows the correctness of the
981: methods used.
982:
983: As is evident from the results, the best accuracy is achieved at values of
984: $x_0$ ($\approx$ the capsule size) about 4--5 metres.
985:
986: %%\Figure{68mm}
987: \Picture{0}{68}{\special{em:graph pic2.gif}}
988: {The SEE method sensitivity to Yukawa forces between
989: the Shepherd and the Particle}
990:
991:
992: \subsection{\nhq Sensitivity to Yukawa forces with $\lambda \sim 1$\,m}
993:
994:
995: In an experiment for finding a Yukawa interaction between the
996: Shepherd and the Particle with the potential (\ref{Yuk2})
997: with $\beta_{1,2}=1$, one computes
998: two theoretical trajectories: one ignoring the
999: Yukawa forces ($x^0(t_i),\ y^0(t_i)$) and another taking
1000: them into account $\Bigl(x^\alpha(t_i),\ y^\alpha(t_i)\Bigr)$. These two
1001: computed curves are compared with the empirical trajectory
1002: using the functional
1003: $S_k$ ($k= 0, \alpha$) according to (\ref{F1}) which may be considered as a
1004: dispersion characterizing a scatter of the ``empirical'' coordinates with
1005: respect to the fitting trajectory. This is true when the theoretical
1006: model is adequate to the real situation. In the case $k=\alpha$ the
1007: functional $S_k= s_\alpha$ has a $\chi^2$ distribution with $n_2=2N-1$
1008: degrees of freedom. With $k=0$ the parameter $\alpha$ is absent,
1009: therefore $S_0$ is distributed according to the $\chi^2$ law with $N_1=2N$
1010: degrees of freedom. Then their ratio $S_0/S_\alpha = F_{n_2,n_1}$ will be
1011: distributed according to the Fischer law with $n_2$ and $n_1$ degrees of
1012: freedom. If an experiment shows that, on a given significance level $q$,
1013: the relation
1014: \beq
1015: S_0/S_\alpha \geq F_{n_1,n_2, q} \label{F5}
1016: \eeq
1017: is valid, one should conclude that a Yukawa force has been
1018: detected. An equality sign shows a minimum detectable force on the given
1019: significance level $q$. We have assumed $q=0.95$.
1020: The results of a sensitivity computation for different
1021: values of the space parameter $\lambda$ are presented in Fig.\,2.%%Table 6.
1022: A maximum sensitivity of $\alpha = 2.1\ten{-7}$ has been observed for
1023: $\lambda= 1.25$ m. This value is 3 to 4 orders of magnitude better than the
1024: sensitivity of terrestrial experiments in the same range.
1025:
1026: %%\Figure{68mm}
1027: \Picture{0}{68}{\special{em:graph pic3.gif}}
1028: {The SEE method sensitivity to Yukawa forces with the range parameter
1029: $\lambda_0$ of the order of the Earth's radius $\RE$}
1030:
1031: These results are based on the measurement method which was proposed in
1032: the original SEE paper \cite{SD92}; as already mentioned, a method
1033: involving an eccentric orbit \cite{Nor}, is much more sensitive and, by
1034: our tentative estimates, can give an error
1035: $\delta\alpha \lsim 10^{-10}$.
1036:
1037: \subsection{Sensitivity to Yukawa forces with $\lambda \sim \RE$}
1038:
1039:
1040: To estimate the parameter $\alpha_0$ in \eqs(\ref{Yuk4}),
1041: computer simulations were carried out using the method
1042: as described above for $\alpha$, based on the Fischer
1043: criterion for the significance level 0.95. The range parameter
1044: $\lambda_0$ varied from $(1/32)\RE$ to $32\RE$.
1045: Two trajectories with the initial
1046: Shepherd-Particle separations $x_0$ of 2 and 5 m were calculated. In
1047: both cases the impact parameter $y_0$ was chosen to be 0.25 m. We used
1048: \eqs (\ref{Yuk4}) with $\alpha=0$, i.e., excluding the non-Newtonian
1049: interaction between the Shepherd and the Particle.
1050: As is evident from \eqs (\ref{Yuk4}), the Particle
1051: trajectory depends on the ratio $r_{01}/\lambda_0$ in the product
1052: $(r_{01}/\lambda_0)\e^{-r_{01}/\lambda_0}$.
1053: This quantity reaches its maximum at $\lambda_0=r_{01}/2$.
1054: Our calculations have confirmed that a maximum
1055: sensitivity of the SEE method ($3.4\ten {-8}$ for $x_0=5$ m)
1056: is indeed observed at this value of $\lambda_0$. This is about an order
1057: of magnitude better than the estimates obtained by other methods.
1058: Hopefully this estimate may be further improved by about an order of
1059: magnitude by optimisation of the orbital parameters.
1060: However, there is a factor which can, to a certain extent, spoil these
1061: results, namely, the uncertainty in the parameter $\omega$ which,
1062: in this calculation, was assumed to be precisely known.
1063:
1064: The simulation results are shown in Fig.\, 3 %Table 7
1065: for two trajectories with initial Shepherd-Particle separations of 2
1066: and 5 metres.
1067:
1068:
1069: \section
1070: {A possible effect of the Earth's radiation belt}
1071:
1072: %%\subsection{The problem}
1073:
1074: Charged particles, penetrating into the SEE capsule from space
1075: and captured by the test bodies, create electrostatic forces that could
1076: substantially distort the experimental results. Among the sources of such
1077: particles one should mention (i) cosmic-ray showers, (ii) solar flares and
1078: (iii) the Earth's radiation belts (Van Allen belts). The effect of
1079: cosmic-ray showers was estimated in Ref. \cite{SD92} and shown to be
1080: negligible. Solar flares are more or less rare events and, although they
1081: create very significant charged particle fluxes, sometimes even exceeding
1082: those in the most dense regions of the radiation belts, one can assume that
1083: the SEE measurements (except those of $\dot G$) are stopped for the period
1084: of an intense flare. On the contrary, the effect of the Van Allen belts is
1085: permanent as long as the satellite orbit passes, at least partially, inside
1086: them.
1087:
1088: We will show here that the charging is unacceptably high at otherwise
1089: favourable satellite orbits, so that some kind of charge removal technique
1090: is necessary, but this problem may be solved rather easily by
1091: presently available technology.
1092:
1093: The range of the most favourable SEE orbital altitudes, roughly 1400 to 3300
1094: km \cite{SD92}, coincides with the inner region of the so-called inner
1095: radiation belt \cite{2}--\cite{5}, situated presumably near the plane
1096: of the magnetic equator. This region is characterized by a
1097: considerable flux of high-energy protons and electrons.
1098: For a SEE satellite at altitudes near 1500 km the duration of the charging
1099: periods is about 12 minutes. Maximum charging rates occur in the central
1100: Atlantic. It should be noted that the South Atlantic Anomaly (SAA) ---
1101: a region of intense Van Allen activity which results from the low
1102: altitude of the Earth's magnetic field lines over the South Atlantic
1103: Ocean --- cannot cause additional problems for the SEE experiments.
1104: The reason is that the SAA mostly contains low-energy protons which cannot
1105: penetrate into the SEE capsule.
1106:
1107: Electrons are known to be stopped by even a thin metallic shell, so
1108: only protons are able to induce charges on the test bodies.
1109: Proton-induced charges on the test bodies can create considerable forces.
1110: The inner radiation belt contains protons with energies of 20 to
1111: 800 MeV, and their maximum fluxes at an altitude of 3000 km over the equator
1112: are as great as about $3\ten 6 \flun$ for energies
1113: $E\gsim 10^6$ eV and about $ 2\ten 4 \flun$ for $E\gsim 10^7$ eV.
1114: At 1500 km altitude these numbers are a few times smaller; the
1115: fluxes gradually decrease with growing latitude $\varphi$ and actually
1116: vanish at $\varphi\sim 40\deg$.
1117:
1118: It is thus necessary to have some estimates taking into account
1119: that (i) the capsule walls have a considerable thickness and stop the
1120: low-energy part of the proton flux and (ii) among the protons that penetrate
1121: the capsule and hit the Particle, the most energetic ones, whose path in the
1122: Particle material is longer than the Particle diameter, fly it through and
1123: hit the capsule wall again. As for the Shepherd, its size is large enough to
1124: stop the overwhelming majority of protons which hit it.
1125:
1126: In what follows, we will assume a Shepherd radius of 20 cm and a Particle
1127: radius of 2 cm and estimate the captured charges for some satellite orbits
1128: in a capsule whose walls of aluminium are 2, 4, 6 and 8 cm thick.
1129: The SEE satellite must actually involve several coaxial cylinders for
1130: thermal-radiation control, and the combined thickness of their walls
1131: must amount to several cm.
1132: We will assume, in addition, that the Particle also consists of aluminium
1133: and stops all protons whose path is shorter than 4 cm (thus a little
1134: overestimating the charge since most of protons will cover a smaller
1135: path through the Particle material). A 100 g Particle of aluminium will have
1136: a radius of $\approx 2.07$ cm.
1137:
1138: It is advisable to determine first which charges (and fluxes that create
1139: them) might be regarded negligible.
1140:
1141:
1142: \subsection {Admissible charges}
1143:
1144:
1145: Let us estimate the Coulomb interaction both between the Shepherd and
1146: the Particle and between each test body and its image in the capsule
1147: walls. To estimate the spurious effects on the
1148: Particle trajectory, it is reasonable to calculate its possible
1149: displacements due to the Coulomb forces from the growing captured
1150: charges. We assume that the test bodies are discharged by grounding
1151: to the capsule before launching the motion.
1152:
1153: \medskip\noi
1154: {\bf Criterion.} We will call the induced charges, or the fields they
1155: create, {\sl admissible} if they cause a displacement of the Particle
1156: with respect to the Shepherd smaller than a prescribed coordinate
1157: measurement error $\delta l$ (we take here $\delta l = 10^{-6}$ m) for
1158: a prescribed measurement time (we take $t \geq 10^4$ s).
1159:
1160: \medskip
1161: A charge on the Shepherd can be estimated as
1162: \beq
1163: q_M \approx eS_M \int J(t,x)\,dt = e S_M F(t,x) \label{B2}
1164: \eeq
1165: where $e$ is the elementary charge,
1166: $x$ is the capsule wall thickness in cm; $J(t,x)$
1167: is the integral proton flux in \flun after passing through the wall,
1168: that is, the flux of protons with energies $E_p > E_p(x)$ where $E_p(x)$
1169: is such an energy that the proton path in aluminium equals $x$ cm; $S_M
1170: \approx 1256$ cm$^2$ is the Shepherd's cross-section; $F(t,x)$ is the
1171: fluence, i.e., the total number of protons of relevant energies that
1172: crosses a square centimeter of area for a certain period $t$.
1173:
1174: In a similar way, the charge captured by the Particle may be found as
1175: \bear
1176: q_m \al\lsim\al eS_m\int [J(t,x)-J(t,\,x{+}4)]\,dt \label{B3} \nn
1177: \eql e S_m \ [F(t,x)-F(t,\,x{+}4)]
1178: \ear
1179: where $S_m \approx 12.56$ cm$^2$ is the Particle cross-section.
1180: The subtraction in the square brackets takes into account the protons
1181: which fly through the Particle without stopping there. The sign
1182: $\lsim$ is used since the effective Particle cross-section is smaller
1183: than its equatorial section.
1184:
1185: The Coulomb acceleration $a_Q (t) = q_M q_m /(r^2 m)$
1186: (in the Gaussian system of units) depends on the Shepherd-Particle
1187: separation $r$ and on the form of the function $J(t)$, which in turn
1188: depends on the satellite orbital motion.
1189:
1190: The charge-induced Particle displacement is approximately
1191: \beq
1192: \Delta l = \int dt \biggl[\int dt\,a_Q(t)\biggr]
1193: \eeq
1194: since the acceleration is almost unidirectional. If, for estimation
1195: purposes, we suppose that the flux is time-independent, $J=J_0=$const,
1196: and take into account that in \eq (\ref{B3}) the difference
1197: $J(t,{x})-J(t,x+4) \approx {2 \over 5}J(t,x)$
1198: (or even smaller; see particular values in the next section), then the
1199: resulting displacement is about
1200: \beq
1201: \Delta l \sim \frac{1}{30}\frac{e^2 S_M S_m J_0^2 t^4}{r^2 m}.\label{B5}
1202: \eeq
1203: The strong time dependence is explained by the
1204: rapid growth of the Coulomb force with capturing the charge. Numerically,
1205: with the above values of $S_M$ and $S_m$, taking $m=100$ g and $r=1$ m
1206: (the latter leads to an overestimated force since the Particle
1207: spends most of time at greater distances), one gets:
1208: \beq
1209: J_0^2 t^4 \lsim 0.83\ten{18}\ {\rm s}^2 {\rm cm}^{-4}. \label{B6}
1210: \eeq
1211: For $t= 10^4$ s an admissible flux is only within 9 \flun.
1212:
1213: Another undesired effect is that the Particle, being charged by the belt
1214: protons, will interact with the capsule walls. This is well
1215: approximated as an interaction with the Particle's mirror image in the
1216: wall, while the latter may be roughly imagined as a conducting plane.
1217: Then, assuming that the Particle is at average at about 25 cm from the
1218: capsule wall and using the same kind of reasoning as above, we obtain
1219: instead of (\ref{B6})
1220: \beq
1221: J_0^2 t^4 \lsim 2.07\ten{19}\ {\rm s}^2 {\rm cm}^{-4} \label{B7}
1222: \eeq
1223: and an admissible proton flux within 45 \flun for $t=10^4$ s.
1224:
1225:
1226: \begin{table*} %\label{T4}
1227: \noi {\bf Table 5.}
1228: Average proton fluxes in some
1229: satellite orbits at minimum solar activity ($x$ is given in cm, $E_p$
1230: in MeV; $i$ is the orbit inclination and $\Omega$ is its ascension
1231: angle, i.e. the longitude at which the satellite crosses the equatorial
1232: plane moving northward. The notation 1.2(3) means $1.2\ten{3}$, etc.
1233: In the first column, the letter `a' labels equatorial orbits, `b' and
1234: `c' mark less and more favourable orbits (thst is, with greater and
1235: smaller numbver of protons), respectively, for given altitide and
1236: inclination.
1237:
1238: \begin{center}
1239: \begin{tabular}{|c|r|r|r|r|r|r|r|r|r|r|}
1240: \hline
1241: & & & & \multicolumn{7}{|c|}{Integral flux for $E_p > E_p(x)$}\\
1242: \cline{5-11}
1243: Orbit & $T$ & $i$ &$\Omega$&$x=0$\z & $ x=2$\z& $x=4$\z & $x=6$\z & $x=8$\z & $x=10$\z & $x=12$\z \\
1244: & s & & &$E_p>0$ & $E_p>65$& $E_p>98$& $E_p>124$& $E_p>146$& $E_p>166$& $E_p>184$\\
1245: \hline
1246: 500b & 5677 & 89\deg & 23.7\deg & 1200 & 8 & 4.5 & 3 & 2.1 & 1.8 & 1.5 \\
1247: 800a & 6053 & 0\deg & & 114 & 66 & 51\zz & 41 & 33\zz & 28\zz & 24\zz \\
1248: 800b & 6053 & 89\deg & 19.4\deg & 2660 & 71 & 46\zz & 34 & 26\zz & 21\zz & 17\zz \\
1249: 1000a & 6307 & 0\deg & & 515 & 325 & 255\zz & 210 & 165\zz & 140\zz & 120\zz \\
1250: 1000b & 6307 & 89\deg & 20\deg\zz & 4120 & 173 & 116\zz & 89 & 65\zz & 54\zz & 44\zz \\
1251: 1000c & 6307 & 89\deg & -83\deg\zz& 51 & 26 & 20\zz & 16 & 12\zz & 10\zz & 8.5 \\
1252: 1500a & 6960 & 0\deg & & 6905 &3160 &2360\zz & 1850 & 1470\zz & 1260\zz & 1070\zz \\
1253: 1500b & 6960 & 102\deg & -12\deg\zz&1.2(4) &1420 &1000\zz & 770 & 600\zz & 500\zz & 415\zz \\
1254: 1500c & 6960 & 102\deg & 97\deg\zz& 2264 &646 &464\zz & 365 & 280\zz & 236\zz & 196\zz \\
1255: 3000a & 9040 & 0\deg & &2.6(5) &1.7(4)&1.15(4) & 9100 & 6500\zz & 5400\zz & 4350\zz \\
1256: 3000b & 9040 & 112\deg & -14.5\deg &1.6(5) &3700 &2400\zz & 1750 & 1300\zz & 1060\zz & 850\zz \\
1257: 3000c & 9040 & 112\deg & 98\deg\zz&9.8(4) &3540 &2350\zz & 1700 & 1300\zz & 1060\zz & 850\zz \\
1258: \hline
1259: \end{tabular}
1260: \end{center}
1261: \end{table*}
1262:
1263: Some more estimates are of interest: if a charge can be kept smaller
1264: than a certain value, then how great may it be to create only negligible
1265: displacements? Suppose that there are constant charges on both the
1266: Shepherd ($q=q_M$) and the Particle ($q=q_m,\ m=100$ g), then they are
1267: admissible according to the above criterion as long as
1268: \bear
1269: q_M q_m \lal <2\ten{-6}\,\qg^2 = \fract{2}{9}\ten{-24}\ {\rm C}^2,
1270: \label{B8}\\
1271: q_m^2 \lal <\half\ten{-6}\,\qg^2. \label{B9}
1272: \ear
1273: These inequalities follow, respectively, from considering
1274: the Shepherd-Particle interaction and the
1275: interaction between the Particle (located at 25 cm from the
1276: wall) and its image. Thus the maximum admissible Particle charge is
1277: about $7\ten{-4}\,\qg \approx 1.5\ten{6} e$; assuming this value, it
1278: follows from (\ref{B8}) that the maximum Shepherd charge is about
1279: $3\ten{-3}\,\qg \approx 5.5 \ten{6} e$. With these
1280: charge values the electric potentials on the test body surfaces are
1281: \bearr
1282: U_M \approx 1.5\ten{-4}\,\qg/{\rm cm} = 45\ {\rm mV}; \nnn
1283: U_m \approx
1284: 3.5\ten{-4}\,\qg/{\rm cm} = 105\ {\rm mV}. \label{b10}
1285: \ear
1286:
1287: If by any means the requirements (\ref{B8}), (\ref{B9}) are satisfied
1288: (e.g., the potentials are kept smaller than the values (\ref{b10})),
1289: the electrostatic effect on the Particle trajectory may be neglected.
1290:
1291: The Shepherd's interaction with its image charge induced in its nearest
1292: bottom of the SEE experimental chamber does not lead to appreciable
1293: Particle displacements. A very demanding requirements on the
1294: Shepherd's charge emerges, however, if the SEE satellite is used for
1295: G-dot determination (whose detailed discussion is postponed to future
1296: papers). One obtains then
1297: \beq
1298: U_M \lsim 12\ {\rm mV}.
1299: \eeq
1300: Evidently, in this case the Shepherd-Particle interaction {\it per se\/}
1301: is not the determining factor with respect to charge limits on the test
1302: bodies.
1303:
1304:
1305: \subsection{Evaluation of charges captured in some orbits}
1306:
1307:
1308: Let us now estimate the charges captured by the Shepherd
1309: and the Particle on board a satellite in various circular orbits for a
1310: single revolution around the Earth, a period of about two hours. Actual
1311: measurement times may exceed this period, but not too much.
1312:
1313: Approximate values of time-averaged proton fluxes are
1314: presented for some circular orbits in Table 5.
1315:
1316: \begin{table*}
1317: {\bf Table 6.}
1318: Average flux, peak flux and captured charges per revolution in some
1319: satellite orbits
1320:
1321: {\small
1322: \begin{center}
1323: \begin{tabular}{|c|c|c|c|r|r|}
1324: \hline
1325: Orbit & Wall & Average flux,& Peak flux,& Shepherd \z & Particle \z \\
1326: & thickness & \flun\z & \flun\z & charge $q_M$ & charge $q_m$ \\
1327: \hline
1328: & 2 cm & \nhq 1420 & \nhq 12300 & 1.5\ten{10} $e$ & 4.5\ten 7 $e$ \\
1329: 1500b & 4 cm & \nhq 1000 & 8800 & 1\ten{10} $e$ & 2.6\ten 7 $e$ \\
1330: & 6 cm & 770 & 6800 & 7.5\ten 9 $e$ & 1.5\ten 7 $e$ \\
1331: & 8 cm & 600 & 5400 & 6\ten 9 $e$ & 1.2\ten 7 $e$ \\
1332: \hline
1333: & 2 cm & 646 & 5700 & 6.5\ten 9 $e$ & 1.9\ten 7 $e$ \\
1334: 1500c & 4 cm & 464 & 4200 & 4.3\ten 9 $e$ & 1.1\ten 7 $e$ \\
1335: & 6 cm & 365 & 3300 & 3.5\ten 9 $e$ & 7\ten 6 $e$ \\
1336: & 8 cm & 280 & 2700 & 2.7\ten 9 $e$ & 5\ten 6 $e$ \\
1337: \hline
1338: \end{tabular}
1339: \end{center}
1340: }
1341: \end{table*}
1342:
1343: The fluxes in Table 5 have been obtained using the computation
1344: software worked out at Nuclear Physics Institute (NPI) of Moscow State
1345: University, called SEE2 (Space Environment Effects 2) and SEREIS
1346: (Space Environment Radiation Effects Information System)
1347: \cite{7}--\cite{9}. This software made use of the NASA models AP8-max and
1348: AP8-min for calculating the proton fluxes \cite{10};
1349: however, the latter rest on measurements performed in
1350: the solar maximum of 1970 and minimum of 1964, while the NPI software
1351: uses some modern models of the Earth's magnetosphere, taking into
1352: account its evolution on the scale of decades.
1353:
1354: The high-energy particle fluxes in the radiation belts are
1355: strongly time-dependent; they vary between maxima and minima of solar
1356: activity, being, at least at low altitudes relevant for a SEE mission,
1357: greater at solar minima [2--5]. Table 5 shows the fluxes at a solar
1358: minimum; similar calculations for a solar maximum show {\it smaller\/}
1359: values by at average 20-25 per cent; the difference exceeds this value
1360: only for $x=0$ (being thus greater for low-energy protons than for
1361: higher-energy ones).
1362:
1363: The solar activity varies from one maximum or minimum to another,
1364: the Earth's magnetic field is sensitive to all these variations and also
1365: varies due to certain terrestrial phenomena. Another source of
1366: uncertainty, probably not a very strong one, is that the shielding effect
1367: is calculated by SEE2 software for a detector placed at the centre of a
1368: spherical shell of shielding material, whereas the capsule is
1369: cylindrical and the angular distribution of the proton flow is also
1370: uncertain. It is thus clear that any values like those presented in
1371: Table 5 may only serve as a guide, giving correct orders of magnitude.
1372:
1373: The above data make it possible to evaluate the captured charges. The
1374: results for two orbits, namely, 1500b and 1500c, are presented in Table
1375: 6. These are the worst and the best variants of orbits at 1500 km among
1376: those analyzed (see the caption of Table 5).
1377:
1378: These and other similar data lead to some conclusions of importance
1379: for the SEE experiments.
1380:
1381: First, the models show zero proton fluxes in equatorial orbits of
1382: 500 --- 800 km altitudes but indicate considerable fluxes at the same
1383: altitudes due to crossing the SAA. It turns out, however, that the SAA
1384: is overwhelmingly a low-energy phenomenon and almost does not affect
1385: fluxes on the relevant energy scale beginning with approximately 65 MeV.
1386: Even more, there is a very small proton flux due to the SAA even at
1387: energies over 10 MeV, so that behind a layer of 1 mm the SAA
1388: influence is already negligible. Therefore, behind a thicker metal layer
1389: there are actually no secondary particles due to SAA protons.
1390:
1391: Second, at 1500 km altitude the fluxes substantially depend on
1392: the orbit orientation but remain on the same scale of a few million
1393: protons per cm$^2$ at energies over 65 MeV.
1394:
1395: Third, evidently, at 3000 km altitude both the total flux (for $x=0$)
1396: and especially its high-energy part are a few times greater than at 1500
1397: km.
1398:
1399: Fourth, and most important: for all orbits in the desirable range of
1400: altitudes the charges are quite large as compared with their admissible
1401: values; they remain large even behind rather thick walls.
1402: It is thus quite necessary to have means to detect and remove the
1403: charges during the measurements.
1404: Moreover, as seen from the peak values in Table 6 and from time
1405: scans of Van Allen charging in orbits of interest (also obtained using
1406: the above-mantioned software), at a charging peak when crossing the
1407: magnetic equator the time required for the charge on the test bodies
1408: to reach its maximum allowable values, as listed above, is a matter of
1409: seconds, not minutes. Therefore the charge must be detected and removed
1410: as it builds up, on a time scale of seconds.
1411:
1412: The detection and measurement of the charge on the test
1413: bodies can probably be achieved relatively easily by an array of minute
1414: microvoltmeters attached to the inner wall of the experimental chamber.
1415:
1416: Several methods for removing positive charge are now being evaluated.
1417: A simple and promising method may be to shoot electron beams directly
1418: at test bodies. The number of electrons needed is on the order of
1419: $10^8/$sec. Although this approach has the inherent drawback that it
1420: requires that an active system must perform correctly for many years, it
1421: is simple in principle and will accomplish the goal.
1422:
1423:
1424: \section{Conclusion}
1425:
1426:
1427: Space offers the prospect of quantum leaps in the accuracy of
1428: gravitational experiments. Although space is a challenging environment
1429: for research, the inherent quiet of space can be exploited to make very
1430: accurate determinations of $G$ and other gravitational parameters,
1431: providing that care is taken to understand the many physical phenomena
1432: in space which have the potential to vitiate accuracy. A distinctive
1433: feature of a SEE mission is its capability to perform such determinations
1434: simultaneously on multiple parameters, making it one of the most promising
1435: proposals.
1436:
1437: To conclude, we enumerate different SEE tests and measurements
1438: and show their expected accuracy as currently estimated:
1439:
1440: \medskip\noi\
1441: \begin{tabular}{ll}
1442: {\it Test/measurement} \cm& {\it Expected accuracy}
1443: \\[8pt]
1444: EP/ISL at a few metres & $2\ten{-7}$ \yy
1445: EP/CD at a few metres & $<10^{-7}\ (\alpha < 10^{-4})$ \yy
1446: EP/ISL at $\sim\RE$ & $<10^{-10}$ \yy
1447: EP/CD at $\sim\RE$ & $<10^{-16}\ (\alpha <10^{-13})$\yy
1448: $G$ & $3.3\ten{-7}$ \yy
1449: $\dot G/G$ & $<10^{-13}$ in one year
1450: \end{tabular}
1451: \medskip
1452:
1453: The last estimate is only tentative; the subject is under study.
1454:
1455: \Acknow
1456: {This work was supported in part by NASA grant \# NAG 8-1442.
1457: K.A.B. wishes to thank Nikolai V. Kuznetsov for helpful discussions and for
1458: providing an access to the SEE2 and SEREIS software.
1459: }
1460:
1461:
1462:
1463: \small
1464: \begin{thebibliography}{99}
1465:
1466: \bibitem{SD92}
1467: A.J. Sanders and W.E. Deeds, Phys. Rev. {\bf D 46}, 480 (1992).
1468:
1469: \bibitem{dSMP}
1470: V. de Sabbata, V.N. Melnikov and P.I. Pronin,
1471: %% ``Theoretical Approach to Treatment of Nonnewtonian Interactions'',
1472: {\it Progr. Theor. Phys. \bf 88}, 623 (1992).
1473:
1474: \bibitem{Mel94}
1475: V.N. Melnikov,
1476: {\it Int. J. Theor. Phys.\/} {\bf 33}, 7, 1569 (1994).
1477:
1478: \bibitem{iztech} % SEE papers 93-98
1479: A.D. Alexeev, K.A. Bronnikov, N.I. Kolosnitsyn, M.Yu.
1480: Konstantinov, V.N. Melnikov and A.G. Radynov,
1481: {\it Izmeritel'naya Tekhnika, } 1993, No. 8, 6; No. 9, 3; No. 10, 6;
1482: 1994, No. 1, 3; {\it Int. J. Mod. Phys. } {\bf D 3}, 4, 773 (1994).
1483:
1484: \bibitem{Gi97}
1485: G.T. Gillies, {\it Rep. Progr. Phys.\/} {\bf 60}, 151 (1997).
1486:
1487: \bibitem{LuTh82}
1488: G.G. Luther and W.R. Towler,
1489: {\it Phys. Rev. Lett.\/} {\bf 48}, 121 (1982).
1490:
1491: \bibitem{FiArm}
1492: M.P. Fitzgerald and T.R. Armstrong,
1493: ``Recent Results for $g$ from MSL Torsion Balance'',
1494: {\it in:\/} ``The Gravitational Constant: Theory and Experiment 200
1495: Years after Cavendish'' (23-24 Nov. 1998), Abstracts
1496: (Cavendish--200 Abstracts), Inst. of Physics, London, 1998.
1497:
1498: \bibitem{Nolting}
1499: J. Schurr, F. Nolting and W. K\"undig,
1500: {\it Phys. Rev. Lett.\/} {\bf 80}, 1142 (1998).
1501:
1502:
1503: \bibitem{Meyer}
1504: H. Meyer,
1505: ``Value for $G$ Using Simple Pendulum Suspensions'',
1506: {\it in:\/} Cavendish--200 Abstracts.
1507: % ``The Gravitational Constant: Theory and Experiment 200
1508: % Years after Cavendish'' (23-24 Nov. 1998), Abstracts, Inst. of Physics,
1509: % London, 1998.
1510:
1511: \bibitem{Karag}
1512: O.V. Karagioz, V.P. Izmaylov and G.T. Gillies,
1513: {\it Grav. \& Cosmol.\/} {\bf 4}, 3, 239 (1998).
1514:
1515: \bibitem{Richman}
1516: S.J. Richman, T.J. Quinn, C.C. Speake and R.S. Davis,
1517: ``Determination of $G$ Using the BIPM Torsion Balance'',
1518: {\it in:\/} Cavendish--200 Abstracts.
1519: % ``The Gravitational Constant: Theory and Experiment 200
1520: % Years after Cavendish'' (23-24 Nov. 1998), Abstracts, Inst. of Physics,
1521: % London, 1998.
1522:
1523: \bibitem{Schwarz}
1524: J.P. Schwarz, J.E. Faller, D.S. Robertson and T.M. Niebauer,
1525: ``The Free-Fall Determination of $G$,
1526: {\it in:\/} Cavendish--200 Abstracts.
1527: % ``The Gravitational Constant: Theory and Experiment 200
1528: % Years after Cavendish'' (23-24 Nov. 1998), Abstracts, Inst. of Physics,
1529: % London, 1998.
1530:
1531: \bibitem{PTB}
1532: %% Experiment by the Physikalish-Technische Bundesanstalt
1533: W. Michaelis,
1534: ``A New Determination of the Gravitational Constant",
1535: {\it Bull. Am. Phys. Soc. \bf 40}, 2, 976 (1995).
1536:
1537: \bibitem{Fis86}
1538: E. Fischbach and C. Talmage,
1539: {\it Mod. Phys. Lett.\/} {\bf A4}, 2303 (1989);
1540: {\it Nature (London)\/} {\bf 356}, 207 (1992).
1541:
1542: \bibitem{Roll64}
1543: P.G. Roll, R. Krotkov and R.H. Dicke,
1544: {\it Ann. Phys. (N.Y.)\/} {\bf 26}, 442 (1964).
1545:
1546: \bibitem{Brag71}
1547: V.B. Braginsky and V.I. Panov, {\it ZhETF} {\bf 61}, 873 (1971)
1548: [{\it Sov. Phys. JETP\/} {\bf 34}, 463 (1972)].
1549:
1550: \bibitem{Adel94}
1551: E.G. Adelberger,
1552: %% "Modern tests of the universality of free fall";
1553: {\it Class. Quantum Grav. \bf 11}, A9--A21 (1994).
1554:
1555: \bibitem{Fuj71-72}
1556: Y. Fujii, {\it Nature Phys. Sci.\/} {\bf 234}, 5 (1971);
1557: {\it Ann. Phys. (N.Y.)\/} {\bf 69}, 494 (1972).
1558:
1559: \bibitem{Long76-84}
1560: D.R. Long, {\it Nature (London)\/} {\bf 356}, 417 (1976);
1561: {\it Nuovo Cim.\/} {\bf 55B}, 252 (1984).
1562:
1563: \bibitem{Adel+}
1564: %% (Adelberger et al. 1987 and 1990, and Su et al., 1994).
1565: E.G. Adelberger et al.,
1566: %% "New Constraints on Composition-Dependent
1567: %% Interactions from the E\"ot-Wash Experiment";
1568: {\it Phys. Rev. Lett. \bf 59}, 8, 849--852 (1987);
1569: %% "Testing the equivalence principle in the field
1570: %% of the Earth: Particle physics at masses below 1 µeV?";
1571: {\it Phys. Rev. \bf D 42}, 10, 3267--3292 (1990).
1572:
1573: \bibitem{Adel97}
1574: E.G. Adelberger, 1997, private communication.
1575:
1576: \bibitem{Ach}
1577: V. Achilli et al.,
1578: %% A geophysical experiment on Newton's inverse-square law.
1579: {\it Nuovo Cim. \bf 112B}, 5, 775 (1997).
1580:
1581: \bibitem{FiG}
1582: E. Fischbach, G.T. Gillies, D.E. Kraus, J.G. Schwan, and C. Talmadge,
1583: %% "Non-Newtonian gravity and new weak forces: an index of measurement
1584: %% and theory";
1585: {\it Metrologia \bf 29}, 3, 13--260 (1992).
1586:
1587: \bibitem{Franklin}
1588: A. Franklin, ``The Rise and Fall of the Fifth Force: Discovery, Pursuit,
1589: and Justification in Modern Physics'', American Institute of Physics,
1590: N.Y., 1993.
1591:
1592: \bibitem{STEP}
1593: J.-P. Blaser et al.,
1594: % M. BYE, G. CAVALLO, T. DAMOUR, C.W.F. EVERITT, A. HEDIN,
1595: % R.W. HELLINGS, Y. JAFRY, R. LAURENCE, M. LEE, A.M. NOBILI, H.J. PAIK,
1596: % R. REINHARD, R. RUMMEL, M.C.W. SANFORD, C. SPEAKE, L. SPENCER,
1597: % P. SWANSON and P.W. WORDEN;
1598: ``STEP (Satellite Test of Equivalence Principle): Report on the
1599: Phase A Study'', ESA/NASA report SCI (93) 4 (March, 1993);
1600: % J.-P. BLASER, A.M. CRUISE, T. DAMOUR, Y. JAFRY, R. LAURENCE, J. LEO'N,
1601: % H.J. PAIK, R. REINHARD, R. RUMMEL and P.W. WORDEN;
1602: ``STEP (Satellite Test of Equivalence Principle):
1603: Assessment Study'', ESA report SCI (94) 5 (May, 1994).
1604:
1605: \bibitem{ZaKol}
1606: N.A. Zaitsev and N.I. Kolosnitsyn. {\it In:\/} ``Experimental Tests of
1607: Gravitation Theory'', Moscow University Press, 1989, p.38--50
1608: (in Russian).
1609:
1610: \bibitem{Nor}
1611: K. Nordtvedt, private communication, 1997.
1612:
1613: \bibitem{2}
1614: B.A. Tverskoy, ``Dynamics of the Earth's Radiation Belts'', Moscow,
1615: Nauka, 1968 (in Russian).
1616:
1617: \bibitem{3}
1618: Yu.I. Galperin, L.S. Gorn and B.I. Khazanov, ``Radiation Measurments in
1619: Space'', Moscow, Atomizdat, 1972 (in Russian).
1620:
1621: \bibitem{4}
1622: %%D.J. Williams, ESSA Technical Report ERL 180-SDL16, 1970.
1623: A.J. Dessler and B.J. O'Brien, ``Penetrating Particle Radiation'',
1624: {\it in:\/} ``Satellite Environment Handbook'', 2nd edition,
1625: ed. F.S. Johnson, Stanford University Press, Stanford, CA, 1965, p.
1626: 53--92.
1627:
1628: \bibitem{5}
1629: W.N. Hess, ``The Radiation Belt and Magnetosphere'', Ginn Blaisdell,
1630: Waltham, MA, 1968.
1631:
1632: \bibitem{7}
1633: V.F. Bashkirov and M.I. Panasiuk,
1634: ``Dynamics of the Trapped Radiation Environment at Low Altitude Region of
1635: the Earth's Magnetosphere".
1636: Workshop ``Space Radiation Environment Modelling: New Phenomena and
1637: Approaches", Oct. 7--9, 1997: Program and Abstracts, p. 1.2.
1638:
1639: \bibitem{8}
1640: V.F. Bashkirov, N.V. Kuznetsov and R.A. Nymmik,
1641: ``Information System for Evaluation of Space Radiation Environment and
1642: Radiation Effects on Spacecraft", ibid., p. 4.8.
1643:
1644: \bibitem{9}
1645: V.F. Bashkirov and M.I. Panasiuk,
1646: {\it Kosmicheskiye Issledovaniya\/} (Space Research), {\bf 36}, 4,
1647: 359--368 (1998).
1648:
1649: \bibitem{10}
1650: D. Bilitza, ``Solar-Terrestrial Models and Application Software",
1651: {\it Planet. Space Sci.\/} {\bf 40}, 4, 541--579 (1992).
1652:
1653: \end{thebibliography}
1654: \end{document}
1655:
1656:
1657: