1: \documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb,prd,floatfix]{revtex4}
2:
3: %\documentclass[prb]{revtex4}% Physical Review B
4:
5: \usepackage{graphicx}% Include figure files
6: \usepackage{dcolumn}% Align table columns on decimal point
7: \usepackage{bm}% bold math
8:
9: %%% HERE ARE OUR SHORTCUTS AND DEFINITIONS
10:
11: \renewcommand{\rho}{\varrho}
12: \newcommand{\ud}{\mbox{d}}
13: \newcommand{\mone}{\multicolumn{1}{c}}
14: \newcommand{\mtwo}{\multicolumn{2}{c}}
15: \newcommand{\mfour}{\multicolumn{4}{c}}
16: \hyphenation{Bun-des-an-stalt}
17:
18:
19: \newcommand{\mum}{\ensuremath \mu\mbox{m}}
20: \newcommand{\mug}{\ensuremath \mu\mbox{g}}
21: \newcommand{\Gunit}{\ensuremath \mbox{m}^3 \;\mbox{kg}^{-1}\;\mbox{s}^{-2}}
22:
23: \newcolumntype{q}{D{.}{.}{1}}
24: \newcolumntype{s}{D{.}{.}{2}}
25: \newcolumntype{t}{D{.}{.}{3}}
26: \newcolumntype{u}{D{.}{.}{4}}
27: \newcolumntype{v}{D{.}{.}{5}}
28: \newcolumntype{w}{D{.}{.}{6}}
29: \newcolumntype{x}{D{.}{.}{7}}
30: \newcolumntype{y}{D{.}{.}{9}}
31:
32:
33:
34:
35: %\documentclass[a4paper,16pt]{article}
36:
37: %\input{header_IR}
38:
39: \begin{document}
40:
41:
42: \title{A Measurement of Newton's Gravitational Constant}
43: \author{St. Schlamminger}
44: \altaffiliation[present address ]{Univ. of Washington, Seattle, Washington, USA}
45: \author{E. Holzschuh}
46: \altaffiliation{deceased}
47: \author{W. K\"undig}
48: \altaffiliation{deceased} \altaffiliation{We dedicate this paper
49: to our colleague Walter K\"{u}ndig, without whose untiring and
50: persistent effort this ambitious experiment would neither have
51: been started nor brought to a successful conclusion. Walter
52: K\"{u}ndig, died unexpectedly and prematurely of a grave illness
53: in May 2005. He conceived the set-up of this experiment and worked
54: on aspects of the analysis until a few days before his death.}
55:
56:
57:
58: \author{F. Nolting}
59: \altaffiliation{Paul Scherrer Institut,Villigen, Switzerland}
60: \author{R.E. Pixley}
61: \email[EMail address: ]{ralph.pixley@freesurf.ch}
62: \author{J. Schurr}
63: \altaffiliation[present address ]{Physikalisch Technische Bundesanstalt, Braunschweig, Germany}
64: \author{U. Straumann}
65: \affiliation{Physik-Institut der Universit\"{a}t Z\"{u}rich, CH-8057 Z\"{u}rich, Switzerland}
66:
67:
68:
69: %---------abstract------------
70: \begin{abstract}
71: A precision measurement of the gravitational constant $G$ has been
72: made using a beam balance. Special attention has been given to
73: determining the calibration, the effect of a possible nonlinearity
74: of the balance and the zero-point variation of the balance. The
75: equipment, the measurements and the analysis are described in
76: detail. The value obtained for $G$ is 6.674252(109)(54) $\times
77: 10^{-11}$\;$\Gunit$. The relative statistical and systematic
78: uncertainties of this result are 16.3$\times 10^{-6}$ and
79: 8.1~$\times 10^{-6}$, respectively.
80: \end{abstract}
81:
82:
83: \pacs{04.80.-y, 06.20.Jr}
84:
85: \maketitle
86:
87:
88:
89:
90: %----------Introduction-----------
91: \section{Introduction}
92: The gravitational constant $G$ has proved to be a very difficult
93: quantity for experimenters to measure accurately. In 1998, the
94: Committee on Data Science and Technology (CODATA) recommended a
95: value of $6.673(10)\times 10^{-11}\; \Gunit$. Surprisingly, the
96: uncertainty, 1,500 parts per million (ppm), had been increased by
97: a factor of 12 over the previously adjusted value of 1986. This
98: was due to the fact that no explanation had been found for the
99: large differences obtained in the presumably more accurate
100: measurements carried out since 1986. Obviously, the differences
101: were due to very large systematic errors. The most recent revision
102: \cite{Mo05} of the CODATA Task Group gives for the 2002
103: recommended value $G=6.6742(10)\times 10^{-11}\; \Gunit$. The
104: uncertainty (150 ppm) has been reduced by a factor of 10 from the
105: 1998 value, but the agreement among the measured values considered
106: in this compilation is still somewhat worse than quoted
107: uncertainties.
108:
109: %-----------------fig 1 principle--------------------
110: \begin{figure}
111: \includegraphics[height=10cm]{prinz.eps}
112: \caption{Principle of the measurement. The FM's are shown in the
113: position together (Pos. T) and the position apart (Pos. A).}
114: \label{fig1}
115: \end{figure}
116:
117:
118:
119: Initial interest in the gravitational constant at our institute
120: had been motivated by reports \cite{Fi86} suggesting the existence
121: of a ``fifth" force which was thought to be important at large
122: distances. This prompted measurements at a Swiss storage lake in
123: which the water level varied by 44~m. The experiment involved
124: weighing two test masses (TM's) suspended next to the lake at
125: different heights. No evidence \cite{Co94,Hu95} was found for the
126: proposed ``fifth" force, but, considering the large distances
127: involved, a reasonably accurate value (750~ppm) was obtained for
128: $G$. It was realized that the same type of measurement could be
129: made in the laboratory with much better accuracy with the lake
130: being replaced by the well defined geometry of a vessel containing
131: a dense liquid such as mercury. Equipment for this purpose was
132: designed and constructed in which two 1.1~kg TM's were alternately
133: weighed in the presence of two moveable field masses (FM's) each
134: with a mass of 7.5~t. A first series of measurements
135: \cite{Sa98,Sb98,No98,No99,Sa99} with this equipment resulted in a
136: value for $G$ with an uncertainty of 220~ppm due primarily to a
137: possible nonlinearity of the balance response function. A second
138: series of measurements was undertaken to eliminate this problem. A
139: brief report of this latter series of measurements has been given
140: in ref. \cite{Sa02} and a more detailed description in a thesis
141: \cite{Sb02}. Since terminating the measurements, the following
142: four years have been spent in improving the analysis and checking
143: for possible systematic errors.
144:
145: Following a brief overview of the experiment, the measurement and
146: the analysis of the data are presented in Secs.~\ref{sec3} entitled
147: Measurement of the Gravitational Signal and Sec.~\ref{sec4} entitled
148: Determination of the Mass-Integration Constant. In Sec.~\ref{sec5}, the
149: present result is discussed and compared with other recent
150: measurements of the gravitational constant.
151:
152:
153:
154:
155: %------------Sec 1, General Considerations----------------
156: \section{General Considerations} The design goal of this
157: experiment was that the uncertainty in the measured value of $G$
158: should be less than about 20~ppm. This is comparable to the quoted
159: accuracy of recent $G$ measurements made with a torsion balance.
160: It is, however, several orders of magnitude better than previous
161: measurements of the gravitational constant (made after 1898)
162: employing a beam balance \cite{Hu95,Mo88,Sp87}.
163:
164: The experimental setup is illustrated in Fig.~\ref{fig1}. Two nearly
165: identical 1.1~kg TM's hanging on long wires are alternately
166: weighed on a beam balance in the presence of the two movable FM's
167: weighing 7.5~t each. The position of the FM's relative to the TM's
168: influence the measured weights. The geometry is such that when the
169: FM are in the position labelled "together", the weight of the
170: upper TM is increased and that of the lower TM is decreased. The
171: opposite change in the TM weights occurs when the FM are in the
172: position labelled "apart". One measures the difference of TM
173: weights first with one position of the FM's and then with the
174: other. The difference between the TM weight differences for the
175: two FM positions is the gravitational signal.
176:
177: The use of two TM's and two FM's has several advantages over a
178: single TM and a single movable FM. Comparing two nearly equivalent
179: TM's tends to cancel slow variations such as zero-point drift of
180: the balance and the effect of tidal variations. Using the
181: difference of the two TM weights doubles the signal. In addition,
182: it causes the influence of the FM motion on the counter weight of
183: the balance to be completely cancelled. Use of two FM's with equal
184: and opposite motion reduces the power required to that of
185: overcoming friction. This also simplified somewhat the mechanical
186: construction.
187:
188: The geometry has been designed such that the TM being weighed is
189: positioned at (or near) an extremum of the vertical force field
190: in both the vertical and horizontal directions for both positions
191: of the FM's. The extremum is a maximum for the vertical position
192: and a minimum for the horizontal position. This double extremum
193: greatly reduces the positional accuracy required in the present
194: experiment.
195:
196: \begin{figure}
197: \includegraphics[width=8.5cm]{aufbau.eps}
198: \caption[A side view of the experiment.]{A side view of the
199: experiment. Legend: 1=measuring room enclosure, 2=thermally
200: insulated chamber, 3=balance, 4=concrete walls of the pit,
201: 5=granite plate, 6=steel girder, 7=vacuum pumps, 8=gear drive,
202: 9=motor, 10=working platform, 11=spindle, 12=steel girder of the
203: main support, 13=upper TM, 14=FM's, 15=lower TM, 16=vacuum tube.}
204: \label{fig2}
205: \end{figure}
206:
207: The measurement took place at the Paul Scherrer Institut (PSI) in
208: Villigen. The apparatus was installed in a pit with thick concrete
209: walls which provided good thermal stability and isolation from
210: vibrations. The arrangement of the equipment is shown in
211: Fig.~\ref{fig2}. The system involving the FM's was supported by a
212: rigid steel structure mounted on the floor of the pit. Steel
213: girders fastened to the walls of the pit supported the balance,
214: the massive (200 kg) granite plate employed to reduce high
215: frequency vibrations and the vacuum system enclosing the balance
216: and the TM's. A vacuum of better than $10^{-4}$~Pa was produced by
217: a turbomolecular pump located at a distance of 2~m from the
218: balance.
219:
220: The pit was divided into an upper and a lower room separated by a
221: working platform 3.5~m above the floor of the pit. All heat
222: producing electrical equipment was located in the upper measuring
223: room. Both rooms had their own separate temperature stabilizing
224: systems. The long term temperature stability in both rooms was
225: better than $0.1^\circ$~C. No one was allowed in either room
226: during the measurements in order to avoid perturbing effects.
227:
228:
229: The equipment was fully automated. Measurements lasting up to 7
230: weeks were essentially unattended. The experiment was controlled
231: from our Zurich office via the internet with data transfer
232: occurring once a day.
233:
234: \section{Measurement of the Gravitational Signal}
235: \label{sec3}
236: We begin this section with a description of the devices employed
237: in determining the gravitational signal. Following the
238: descriptions of these devices, the detailed schedule of the
239: various weighings and their analysis are given. Balance weighings
240: will be expressed in mass units rather than force units. The value
241: of local gravity was determined for us by E. E. Klingel\'{e} of
242: the geology department of the Swiss Federal Institute. The
243: measurement was made near the balance on Sept. 11, 1996 using a
244: commercial gravimeter (model G \#317 made by the company
245: LaCoste-Romberg). The value found was $9.807 233 5(6)\;\mbox{m}\mbox{s}^{-2}$.
246: This value was used to convert the balance readings into force
247: units.
248:
249: %---------------sec 3.1--------------------
250: \subsection{The Balance}
251: \label{sec31} The beam balance was a modified commercial mass
252: comparator of the type AT1006 produced by the Mettler-Toledo
253: company. The mass being measured is compensated by a counter
254: weight and a small magnetic force between a permanent magnet and
255: the current flowing in a coil mounted on the balance arm. An
256: opto-electrical feedback system controlling the coil current
257: maintains the balance arm in essentially a fixed position
258: independent of the mass being weighed. The digitized coil current
259: is used as the output reading of the balance.
260:
261: The balance arm is supported by two flexure strips which act as
262: the pivot. The pan of the balance is supported by a parallelogram
263: guide attached to the balance frame. This guide allows only
264: vertical motion of the pan to be coupled to the arm of the
265: balance. Horizontal forces produced by the load are transmitted to
266: the frame and have almost no influence on the arm.
267:
268: As supplied by the manufacturer, the balance had a measuring range
269: of 24~g above the 1~kg offset determined by the counter weight.
270: The original readout resolution was 1~\mug\ and the specified
271: reproducibility was 2~\mug. The balance was designed especially
272: for weighing a 1 kg standard mass such as is maintained in many
273: national metrology institutes.
274:
275: In the present experiment, the balance was modified by removing
276: some nonessential parts of the balance pan which resulted in its
277: weighing range being centered on 1.1~kg instead of the 1~kg of a
278: standard mass. Therefore, 1.1~kg TM's were employed. In order to
279: obtain higher sensitivity required for measuring the approximately
280: 0.8~mg difference between TM weighings, the number of turns on the
281: coil was reduced by a factor of 6, thus reducing the range to 4~g
282: for the same maximum coil current. The balance was operated at an
283: output value near 0.6~g which gave a good signal-to-noise ratio
284: with low internal heating. For the present measurements, a mass
285: range of only 0.2~g was required. The full readout resolution of
286: the analog to digital converter (ADC) measuring the coil current
287: was employed which resulted in a readout-mass resolution of
288: 12.5~ng.
289:
290: An 8th order low-pass, digital filter with various time constants
291: was available on the balance. Due to the many weighings required
292: by the procedure employed to cancel nonlinearity (see
293: Sec.~\ref{sec34}), it was advantageous to make the time taken for
294: each weighing as short as possible. Therefore, the shortest filter
295: time constant (approximately 7.8~s) was employed and output
296: readings were taken at the maximum repetition rate allowed by the
297: balance (about 0.38~s between readings).
298:
299: Pendulum oscillations were excited by the TM exchanges. Small
300: oscillation amplitudes (less than 0.2~mm) of the TM's
301: corresponding to one and two times the frequency for pendulum
302: oscillations (approximately 0.26~Hz for the lower TM and 0.33~Hz
303: for the upper TM's) were observed. They were essentially undamped
304: with decay times of several days. The unwanted output amplitudes
305: of these pendulum oscillation were not strongly attenuated by the
306: filter (half-power frequency of 0.13~Hz) and therefore had to be
307: taken into account in determining the equilibrium value of a
308: weighing.
309:
310:
311: The equilibrium value of a weighing was determined in an on-line,
312: 5-parameter, linear least-squares fit made to 103 consecutive
313: readings of the balance starting 40~s after a load change. The
314: parameters of the fit were 2 sine amplitudes, 2 cosine amplitudes
315: and the average weight. The pendulum frequencies were known from
316: other measurements and were not parameters of the on-line fit. The
317: 40~s delay before beginning data taking was required in order to
318: allow the balance to reach its equilibrium value (except for
319: oscillations) after a load change. This procedure (including the
320: 40~s wait) is what we call a "weighing". A weighing thus required
321: about 80~s.
322:
323: Data of a typical weighing and the fit function used to describe
324: their time distribution are shown in the upper part of
325: Fig.~\ref{fig3}. The residuals $\delta$ divided by a normalization
326: constant $\sigma$ are shown in the lower part of this figure. The
327: normalization constant has been chosen such that the rms value of
328: the residuals is 1. Since the balance readings are correlated due
329: to the action of the digital filter, the value of $\sigma$ does
330: not represent the uncertainty of the readings. It is seen that the
331: residuals show only rather wide peaks. These peaks are probably
332: due to very short random bursts of electronic noise which have
333: been widened by the digital filter. With the sensitivity of our
334: modified balance, they represent a sizable contribution to the
335: statistical variations of the weighings. They are of no importance
336: for the normal use of the AT1006 balance.
337: %--------------fig 3 (103 readings of a weighing)--------------
338: \begin{figure}
339: \includegraphics[width=8.5cm]{pendulum_residual.ps}
340: \caption{Shown in the upper plot are the balance readings for a
341: typical weighing illustrating the oscillatory signal due to
342: pendulum oscillations. The output is the uncalibrated balance
343: reading corresponding to approximately 1.1 kg with a magnetic
344: compensation of 0.6 g. The amplitude of the oscillatory signal
345: corresponds to about 1.5 \mug. The lower plot shows the normalized
346: residuals. The normalization has been chosen such that their rms
347: value is 1.} \label{fig3}
348: \end{figure}
349:
350:
351:
352: A direct calibration of the balance in the range of the 780 \mug\
353: gravitation signal can not be made with the accuracy required
354: in the present experiment ($< 20$~ppm) since calibration masses of
355: this size are not available with an absolute accuracy of better
356: than about 300~ppm. Instead, we have employed a method in which an
357: accurate, coarse grain calibration was made using two 0.1 g
358: calibration masses (CM's). The CM's were each known with an
359: absolute accuracy of 4~ppm. A number of auxiliary masses (AM's)
360: having approximate weights of either 783 $\mu g$ or $16\times
361: 783=12,528$~\mug\ were weighed along with each TM in steps of
362: 783~\mug\ covering the 0.2~g range of the CM. Although the AM's
363: were known with an absolute accuracy of only 800~ng (relative
364: uncertainty 1,000~ppm), the method allowed balance nonlinearity
365: effects to be almost entirely cancelled. Thus, the effective
366: calibration accuracy for the average of the TM difference
367: measurements was essentially that of the CM's. A detailed
368: description of this method is given in Sec.~\ref{sec310}.
369:
370:
371:
372: In our measurements, the balance was operated in vacuum. The
373: balance proved to be extremely temperature sensitive which was
374: exacerbated by the lack of convection cooling in vacuum. The
375: measured zero-point drift was 5.5~mg/$^\circ \mbox{C}$. The sensitivity
376: of the balance changed by 220~ppm$/^\circ \mbox{C}$. To reduce these
377: effects, the air temperature of the room was stabilized to about
378: 0.1~$^\circ \mbox{C}$. A second stabilized region near the balance was
379: maintained at a constant temperature to 0.01~$^\circ \mbox{C}$. Inside
380: the vacuum, the balance was surrounded by a massive (45~kg) copper
381: box which resulted in a temperature stability of about 1~mK.
382: Although zero-point drift under constant load for a 1~mK
383: temperature change was only 5.5~\mug\ , the effects of self
384: heating of the balance due to load changes during the measurement
385: of the gravitational signal were much larger. Details of this
386: effect and how they were corrected are described in Sec.\ref{sec37}.
387:
388:
389: \subsection{The Test Masses}
390: \label{sec32} One series of measurements was made using copper
391: TM's and two with tantalum TM's. Various problems with the mass
392: handler occurred during the measurements with the tantalum TM's
393: which resulted in large systematic errors. Although the tantalum
394: results were consistent with the measurements with the copper
395: TM's, the large systematic errors resulted in large total errors.
396: The tantalum measurements were included in our first publication,
397: but we now believe that better accuracy is obtained overall with
398: the copper measurements alone. We therefore describe only the
399: measurements made with the copper TM's in the present work.
400:
401: A drawing of a copper TM is shown in Fig.~\ref{fig:cotm}. The
402: 45~mm diameter, 77 mm high copper cylinders were plated with a
403: 10~\mum\ gold layer to avoid oxidation. The gold plating was made
404: without the use of nickel in order to avoid magnetic effects. Near
405: the top of each TM on opposite sides of the cylinder were two
406: short horizontal posts. The posts were made of Cu-Be (Berylco 25).
407: The tungsten wires used to attach the TM's to the balance were
408: looped around these posts in grooves provided for this purpose.
409: The wires had a diameter of 0.1~mm and lengths of 2.3~m for the
410: upper TM and 3.7~m for the lower. The loop was made by crimping
411: the tungsten wire together in a thin copper tube. A thin,
412: accurately machined, copper washer was placed in a cylindrical
413: indentation on the top surface of the lower TM in order to trim
414: its weight (including suspension) to within about 400~\mug\ of
415: that of the upper TM and suspension.
416:
417: %-------------fig 4 schematic of TM------------------
418: \begin{figure}
419: \includegraphics[width=8.5cm]{nn_TM_drawing.eps}
420: \caption{Drawing of TM inside the vacuum tube. Dimensions are
421: given in mm. } \label{fig:cotm}
422: \end{figure}
423:
424: Measurement of the TM's dimensions was made with an accuracy of
425: 5~\mum\ using the coordinate measuring machine (CMM) at PSI. The
426: weight of the gold plating was determined from the specified
427: thickness of the layer. The weight of the tungsten wires was
428: determined from the dimensions and the density of tungsten. The
429: thin tubing used to crimp the tungsten wires was weighed directly.
430: The weight of the complete TM's was determined at the
431: Mettler-Toledo laboratory with an accuracy of 25~\mug\ (0.022~ppm)
432: before and after the gravitational measurement. It was found that
433: the mass of both TM's had increased by a negligible amount
434: (0.5~ppm) during the measurement.
435:
436: An estimate of possible density gradients in the TM's was
437: determined by measuring the density of copper samples bordering
438: the material used for making the TM's. It was found that the
439: variation of the relative density gradients over the dimensions of
440: either TM was less than $2\times 10^{-4}$ in both the longitudinal
441: and the radial directions.
442:
443: \subsection{TM Exchanger}
444: \label{sec33} In weighing the TM's, it was necessary to remove the
445: suspension supporting one TM from the balance and replace it by
446: the other supporting the other TM. The exchange was accomplished
447: by a step-motor driven hydraulic systems to raise the suspension
448: of one TM while lowering the other. A piezo-electric transducer
449: mounted above the pan of the balance was used to keep the load on
450: the balance during the exchange as constant as possible. This was
451: done in order to avoid excessive heating due to the coil current
452: and to reduce anelastic effects in the flexure strips supporting
453: the balance arm. The output excursions were typically less than
454: 0.1~g. The exchange of the TM's required about 4~min.
455:
456: The TM suspension rested on a thin metal arm designed to bend
457: through 0.6~mm when loaded with 1.1~kg. Therefore, the transfer of
458: TM's was accomplished with a vertical movement of typically 2~mm
459: (0.6~mm bending of the spring plus an additional 1.4~mm to avoid
460: electrostatic forces). The metal arm was attached to a
461: parallelogram guide (similar to that of the balance) to assure
462: only vertical motion.
463:
464: Although the parallelogram reduced the error resulting from the
465: positioning of the load, it was nevertheless important to have the
466: TM load always suspended from the same point on the balance pan.
467: This was accomplished by means of a kinematic coupling
468: \cite{Sl88,Fu81}. The coupling consisted of three pointed titanium
469: pins attached to each TM suspension which would come to rest in
470: three titanium V grooves mounted on the balance pan. The
471: reproducibility of this positioning was 10~\mum. The pieces of the
472: coupling were coated with tungsten carbide to avoid electrical
473: charging and reduce friction.
474:
475:
476:
477:
478: \subsection{Auxiliary Masses}
479: \label{sec34} In order to correct for any nonlinearity of the
480: balance in the range of the signal, use was made of many auxiliary
481: masses (AM's) spanning the 200~mg range of the CM's in steps of
482: approximately 783~\mug. Although the AM's could not be measured
483: with sufficient accuracy to calibrate the balance absolutely, they
484: were accurate enough to correct the measured gravitational signal
485: for a possible nonlinearity of the balance. Each TM was weighed
486: along with various combinations of AM's. One essentially averaged
487: the nonlinearity over the 200 mg range of the CM's in 256 load
488: steps of 200~mg/256=783~\mug. A weighing of both 100~mg CM's was
489: then used to determine the absolute calibration of the balance
490: which is valid for the TM weighings averaged over this range. The
491: effect of any nonlinearity essentially cancels due to the
492: averaging process. The accuracy of the nonlinearity correction is
493: described in Sec.~\ref{sec310}.
494:
495: The 256 load steps were accomplished using 15 AM's with a mass of
496: approximately 783~\mug\ called AM1's and 15 AM's with 16 times
497: this mass (12,528~\mug) called AM16's. They were made from short
498: pieces of stainless steel wire with diameters of 0.1 mm and
499: 0.3~mm. The wires were bent through about $70^\circ$ on both ends
500: leaving a straight middle section of about 6~mm. The mass of the
501: AM's were electrochemically etched to obtain as closely as
502: possible the desired masses. The RMS deviation was 1.5~\mug\ for
503: the AM1's and 2.3~\mug\ for the AM16's.
504:
505: By weighing a TM together with various AM combinations, one
506: obtains the value of the TM weight simultaneously with the
507: linearity information. The only additional time required for this
508: procedure over that of weighing only the TM's is the time
509: necessary to change an AM combination (10 to 30~s).
510:
511:
512:
513: \subsection{Mass Handler}
514: \label{sec35} The mass handler is the device which placed the AM's
515: and the CM's on the balance or removed them from the balance. The
516: mass handler was designed by the firm Metrotec AG. The operation
517: of this device is illustrated in the somewhat simplified drawing
518: of Fig.~\ref{fig:masshandler} showing how the AM1's and the CM1
519: are placed on the metal strip attached to the balance pan. Only 6
520: of the 14 steps are shown in this illustration for clarity. The
521: portion of the handler used for the AM'16 and the CM2 (not shown)
522: is similar except that the AM'16 are placed on a metal strip
523: located below the one used for the AM1's. All of the AM1's
524: pictured in Fig.~\ref{fig:masshandler} are lying on the steps of a
525: pair of parallel double staircases. The staircases are separated
526: by 6~mm which is the width of the AM's between the bent regions on
527: both ends. The spacing between the staircases is such that they
528: could pass on either side of the horizontal metal strip fastened
529: to the balance pan as the staircases were moved up or down. The
530: motion of each staircase pair was constrained to the vertical
531: direction by a parallelogram (similar to those of the balance)
532: fastened to the frame of the mass handler. The staircases for
533: AM1's and AM16's were moved by two separate step motors located
534: outside the vacuum system. The step motors were surrounded by mu
535: metal shielding to reduce the magnetic field in the neighborhood
536: of the balance. Moving the staircases down deposited one AM after
537: another onto the metal strip. Moving the staircases up removed the
538: AM's lying on the strip. The steps of the staircase had hand
539: filed, saddle shaped indentations to facilitate the positioning of
540: the AM's. The heights of the steps were 2~mm and the steps on the
541: left side of the double staircase were displaced in height by 1~mm
542: from those on the right. Thus, the AM1's were alternately placed
543: on the balance to the left and to the right of the center of the
544: main pan in order to minimize the torque which they produced on
545: the balance.
546: %----------------fig 5-------------------
547: \begin{figure}
548: \includegraphics[width=8cm]{masshandler.eps}
549: \caption{Simplified drawing of the mass handler illustrating the
550: principle of operation. Legend: 1=pivoted lever pair holding a CM,
551: 2=narrow strip to receive the CM, 3=double stair case pair holding
552: AM's, 4=narrow strip to receive AM's, 5=balance pan, 6 flat
553: spring, 7=frame, 8=stepmotor driven cogwheel, and 9=coil spring.
554: The pivoted-lever pair and the double-staircase pair are spaced
555: such that they can pass on either side of the narrow strips 2 and
556: 4 fastened to the balance pan. The two flat springs 6 form two
557: sides of a parallelogram which assures vertical motion of the
558: double stair case pair.} \label{fig:masshandler}
559: \end{figure}
560:
561: Raising the staircase structure above the position shown in the
562: figure caused a rod to push against a pivoted lever holding CM1.
563: With this operation, CM1 was placed on the upper strip attached to
564: the balance. Reversing the operation allowed the spring to move
565: the lever in the opposite direction and remove CM1 from the
566: balance.
567:
568: Due to the very small mass of the AM1's, difficulty was
569: occasionally experienced with the AM1's sticking to one side of
570: the staircase or the other. The staircases were made of aluminum
571: and were coated with a conductive layer of tungsten carbide to
572: reduce the sticking probability. Sticking nevertheless did occur.
573: The sticking would cause an AM1 to rest partly on the staircase
574: and partly on the pan, thus giving a false balance reading. In
575: extreme cases, the AM1 would fall from the holder and therefore be
576: lost for the rest of the measurement. No problem was experienced
577: with the heavier AM16's and the CM's.
578:
579:
580: \subsection{Weighing Schedule}
581: \label{sec36} The experiment was planned so that the zero-point
582: (ZP) drift and the linearity of the balance could be determined
583: while weighing a TM. In principle one needs just 4 weighings
584: (upper and lower TM with FM's together and apart) to determine the
585: signal for each AM placed on the balance. Repeating these 4
586: weighings allows one to determine how much the zero point has
587: changed and thereby correct for the drift. Since there are 256 AM
588: values required to correct the nonlinearity of the balance, a
589: minimum of 2048 weighings is needed for a complete determination
590: of the signal corrected for ZP drift and linearity. One also
591: wishes to make a number of calibration measurements during the
592: series of measurements.
593:
594: The order in which the measurements are performed influences
595: greatly the ZP drift correction of measurements. Changing AM's
596: requires only 6 to 30~s, while exchanging TM on the balance takes
597: about 230 s and moving the FM from one position to the other
598: requires about 600 s. These times are to be compared with the 80 s
599: required for a weighing and about 1 hr for a complete calibration
600: measurement (see Sec.~\ref{sec39}). One therefore wishes to
601: measure a number of AM values before exchanging TM, and repeat
602: these measurements for the other TM before changing the FM
603: positions or making a calibration.
604:
605: The schedule of weighing adopted is based on several basic series
606: for the weighing of the different TM's with different FM
607: positions. The series are defined as follows:
608: \begin{enumerate}
609: \item An S4 series is defined as the weighing of four successive
610: AM values with a particular TM and with all weighing made for the
611: same FM positions. \item An S12 series involves three S4 series
612: all with the same four AM values and the same FM positions. The S4
613: series are measured first for one TM, then the other TM and
614: finally with the original TM. \item An S96 is eight S12 series,
615: all made with the same FM positions and with the AM values
616: incremented by four units between each S12 series. A TM exchange
617: is also made between each S12 series. An S96 series represents the
618: weighings with 32 successive AM values for both TM all with the
619: same FM positions. \item An S288 series is three S96 series, first
620: with one FM position, then the other and finally with the original
621: FM position. A calibration measurement is made at the beginning of
622: each S288 series. Thus, the S288 series represents the weighings
623: with 32 successive AM values for both TM's and both FM positions
624: and includes its own calibration. \item An S2304 series is made up
625: of eight S288 series with the AM values incremented by 32 between
626: each S288 series. An S2304 series completes the full 256 AM values
627: with weighings of both TM's and both FM positions.
628: \end{enumerate}
629:
630: A total of eight valid S2304 series was made over a period of 43
631: days. Alternate S2304 series were intended to be made with
632: increasing and decreasing AM values. Unfortunately, the restart
633: after a malfunction of the temperature stabilization in the
634: measuring room was made with the wrong incrementing sign. This
635: resulted in five S2304 series being made with increasing AM values
636: and three with decreasing.
637:
638: \subsection{Analysis of the Weighings}
639: \label{sec37} In ref.~\cite{Sa02,Sb02}, the so called ABA method
640: was used to analyze the data obtained from the balance and thereby
641: obtain the difference between the mass of the A and B TM. This
642: method assumes a linear time dependence of the weight that would
643: be obtained for the A TM at the time when the B TM was measured
644: based on the weights measured for A at an earlier and a later
645: time. However, a careful examination of the data showed that the
646: curvature of the ZP drift was quite large and was influenced by
647: the previous load history of the balance. This indicated that the
648: linear approximation was not a particularly good approximation. We
649: have therefore reanalyzed the data using a fitting procedure to
650: determine a continuous ZP function of time for each S96 series.
651: The data and fit function for a typical S96 series starting with a
652: calibration measurement is shown in Fig.~\ref{fig:zeropoint}. The
653: procedure used to determine the ZP data and fit are described in
654: the following. The criterion for a valid weighing is described in
655: Sec.~\ref{sec38}.
656:
657: The data of Fig.~\ref{fig:zeropoint} show a slow rise during the
658: first hour after the calibration measurement followed by a
659: continuous decrease with a time constant of several hours. These
660: slow variations are attributed to thermal variations resulting
661: principally from the different loading of the balance during the
662: calibration measurement. Superposed on the slow variations are
663: rapid variations which are synchronous with the exchange of the
664: TM's. The rapid variations peak immediately after the TM exchange
665: and decrease thereafter with a typical slope of 0.3 ~\mug\ /hr.
666: The cause of the rapid variations is unknown.
667:
668: The data employed in the ZP determination were the weighings of
669: the upper and lower TM's for the S96 series. The known AM load for
670: each weighing was first subtracted to obtain a net weight for
671: either TM plus the unknown zero-point function at the time of each
672: weighing. A series of Legendre polynomials was used to describe
673: the slow variations of the zero-point function. A separate $P_0$
674: coefficient was employed for each TM. The rapid variations were
675: described by a sawtooth function starting at the time of each TM
676: exchange. The fit parameters were the coefficient of the Legendre
677: polynomials and the amplitude of the sawtooth function. The
678: sawtooth amplitude was assumed to be the same for all rapid peaks
679: of an S96 series. The sawtooth function was used principally to
680: reduce the $\chi^2$ of the fit and had almost no effect on the
681: results obtained when using the ZP function. All parameters are
682: linear parameters so that no iteration is required. The actual ZP
683: function is the sawtooth function and the polynomial series
684: exclusive of the time independent terms (i.e. the sum of the
685: coefficients times $P_n(0)$ for even $n$).
686:
687: %------------fig 6 zero-point variation---------------
688: \begin{figure}
689: \includegraphics[width=8.5cm]{zp_residual.ps}
690: \caption{The zero-point variation as a function of time for a
691: typical S96 series including calibration is shown in the upper
692: part of this figure. The solid curve is the fit function starting
693: after the last dummy weighing. The fit function for this S96
694: series has 76 degrees of freedom. The normalized residuals
695: $\delta/\sigma$ are shown in the lower plot. The normalization of
696: the residuals has been chosen such that their rms value is 1.}
697: \label{fig:zeropoint}
698: \end{figure}
699:
700: Such calculations were made for various numbers of Legendre
701: coefficients in the ZP function. It was found that the
702: gravitational signal was essentially constant for a maximum order
703: of Legendre polynomials between 8 and 36. In this range of
704: polynomials, the minimum calculated signal was 784.8976(91)~\mug\
705: for a maximum order equal to 22 and a maximum signal of
706: 784.9025(93)~\mug\ for a maximum order equal to 36 (i.e. a very
707: small difference). In all following results, we shall use the
708: signal 784.8994(91)~\mug\ obtained with a maximum polynomial order
709: of 15.
710:
711: It has been implicitly assumed in the above ZP determination that
712: the AM load values were known with much better accuracy than the
713: reproducibility of the balance producing the data used in the ZP
714: fit. Although the AM values were sufficiently accurate for
715: determining the general shape of the ZP function, their relative
716: uncertainties were comparable to the uncertainties of the balance
717: data used in the fit. The $P_0$ mass parameters of the TM's
718: obtained from the fit were therefore not used for the TM
719: differences at the two FM positions which are needed in order to
720: determine the gravitational signal. Instead, the value of the ZP
721: function was subtracted from each weighing, and an ABA mass
722: difference was determined for each triplet of weighings having the
723: same AM load. Since the mass of the AM's do not occur in this TM
724: difference, they do not influence the calculation. The ABA
725: calculation is valid for this purpose since the ZP corrected
726: weighings have essentially no curvature. Such TM differences
727: determined for the apart-together positions of the FM's are then
728: used to calculate the gravitational signal.
729:
730:
731:
732: The TM differences for the apart-together positions of the FM's as
733: a function of time are shown for a ZP corrected S288 series in
734: Fig.~\ref{fig:S288}. Individual data points are resolved in the
735: magnified insert of this figure. Each data point is the B member
736: of a TM difference obtained from an ABA triplet in which all
737: weighings have the same load value.
738:
739: %--------------fig 7-------------------
740: \begin{figure}
741: \includegraphics[width=8.5cm]{S288.ps}
742: \caption{The measured weight difference in $\mu g$ between TM's
743: obtained from an S288 series. The magnified insert shows the
744: individual TM differences which are not resolved in the main part
745: of the figure.} \label{fig:S288}
746: \end{figure}
747:
748: All of the TM differences (ZP corrected) for the entire experiment
749: are shown in Fig.~\ref{fig:tmdiff}. The data labelled apart have
750: been shifted by 782~\mug\ in order to allow both data sets to be
751: presented in the same figure. A slow variation of 2.5~\mug\ in
752: both TM differences occurred during the 43 day measurement. Also
753: seen in this figure is a 0.7~\mug\ jump which occurred in the data
754: for both the apart and together positions of the FM's on day 222.
755: The slow variation is probably due to sorption-effect differences
756: of the upper and lower TM's. The jump was caused by the loss or
757: gain of a small particle such as a dust particle by one of the
758: TM's. In order to determine the gravitational signal, an ABA
759: difference was calculated for apart-together values having the
760: same AM load. The slow variation seen in Fig.~\ref{fig:tmdiff} is
761: sufficiently linear so that essentially no error results from the
762: use of the ABA method. The jump in the apart-together differences
763: caused no variation of the gravitational signal.
764:
765:
766: %------------fig 8 plot of TM difference vs time---------------
767: \begin{figure}
768: \includegraphics[width=8.5cm]{signal_both.ps}
769: \caption{The measured weight difference in $\mu g$ between TM's
770: for the FM's positions apart and together. The measured values for
771: the FM apart have been displaced 782 $\mu g$ in order to show both
772: types of data in the same figure. } \label{fig:tmdiff}
773: \end{figure}
774:
775: In Fig.~\ref{fig:histo} is shown a plot of the binned difference
776: between the FM apart-together positions for all valid data (see
777: Sec.~\ref{sec38}. The differences were determined using the ABA
778: method applied to weighings made with the same AM loads. Also
779: shown in the figure is a Gaussian function fit to the data. The
780: data are seen to agree well with the Gaussian shape which is a
781: good test for the quality of experimental data. The
782: root-mean-square (RMS) width of the data is 1.03 time the width of
783: the Gaussian function. The true resolution for these weighings may
784: be somewhat different than shown in Fig.~\ref{fig:histo} due to
785: the fact that the data have not been corrected for nonlinearity of
786: the balance (see Sec.~\ref{sec310}) and for correlations due to
787: the common ZP function . Nevertheless, these effects would not be
788: expected to influence the general Gaussian form of the
789: distribution.
790:
791: %------------------fig 9 gauss resolution------------
792: \begin{figure}
793: \includegraphics[width=8.5cm]{cu_bin.ps}
794: \caption{Binned data for the FM apart-together weight differences
795: (points) and a fitted Gaussian function (curve) shown as a
796: deviation from the mean difference. Poisson statistics were used
797: to determine the uncertainties.} \label{fig:histo}
798: \end{figure}
799:
800:
801:
802: A plot of the signal obtained for the S2304 series with increasing
803: and decreasing load is shown in Fig.~\ref{fig:cycles}. The average
804: signal for increasing load is 784.9121(125)~\mug\ and the average
805: for decreasing load is 784.8850(133)~\mug. The common average for
806: both is 784.8994(91)~\mug. The averages for increasing load and
807: for decreasing load lie within the uncertainty of the combined
808: average. This shows that the direction of load incrementing did
809: not appreciably influence the result.
810:
811: %------------fig 10 cycles plot---------------
812: \begin{figure}
813: \includegraphics[width=8.5cm]{cycles_new.ps}
814: \caption{The average signal for each of the eight S2304 series.
815: Series with increasing load are shown as circles. Series with
816: decreasing load are shown as squares. The dashed line is the
817: average of all eight series.} \label{fig:cycles}
818: \end{figure}
819:
820:
821:
822:
823:
824: Although the weighings making up an S96 series are correlated due
825: to the common ZP function determined for each S96 series, the
826: results of each S96 series, in particular the TM parameter, are
827: independent. The 32 signal values obtained from the three S96
828: series making each S288 series are also independent. However,
829: since the nonlinearity correction (see Sec.~\ref{sec310}) being
830: employed is applicable only to an entire S2304 series (not to
831: individual S288 series), it is only the eight S2304 series which
832: should be compared with one another. This restricts the way in
833: which the average signal is to be calculated for the entire
834: measurement, namely the way in which the data are to be weighted.
835:
836: We have investigated two weighting procedures. In the first, each
837: S2304 series average was weighted by the number of valid triplets
838: in that series. This assumes that the weighings measured in all
839: S2304 series have the same a priori accuracy. In the second
840: method, it was assumed that the accuracy for each weighing in a
841: series was the same but might be different for different series.
842: We believe the second method is the better method since it takes
843: into account changes that occur during the long, 43 day
844: measurement (e.g. the not completely compensated effects of
845: vibration, tidal forces and temperature). The averages obtained
846: with the two methods differ by approximately 6~ng with the second
847: method giving the smaller average signal. This is a rather large
848: effect. It is only slightly smaller than the statistical
849: uncertainty of 9 to 10~ng obtained for either method. In the rest
850: of this work, we shall discus only the results obtained with the
851: second method.
852:
853:
854: \subsection{Criterion for Valid Data}
855: \label{sec38}
856: Two tests were used to determine whether a measured weighing was
857: valid. An on-line test checked whether the $\chi^2$ value of the
858: fit to the pendulum oscillations was reasonable. A large value
859: caused a repeat of the weighing. After two repeats with large
860: $\chi^2$, the measurement for this AM value was aborted. An
861: aborted weighing usually indicated that the AM was resting on the
862: mass handler and on the balance pan in an unstable way.
863:
864: A more frequent occurrence was that of an AM which rested on both
865: the mass handler and the balance was almost stable thereby giving
866: a reasonable $\chi^2$. In order to reject such weighings, an
867: off-line calculation was made to check whether the measured weight
868: was within 10~\mug\ of the expected weight. The statistical noise
869: of a valid weighing was typically about 0.15~\mug\ (see
870: Fig.~\ref{fig:histo} showing ABA difference involving 3
871: weighings). Excursions of more than 10~\mug\ were thus a clear
872: indication of a malfunction.
873:
874: This off-line test is somewhat more restrictive than the off-line
875: test employed in our original analysis. In the original analysis,
876: a check was made only to see that the weight difference between
877: the TM's for equal AM loadings was reasonable. The more
878: restrictive test used in the present analysis resulted in the
879: rejection of the S2304 series at the time when the room
880: temperature stabilizer was just beginning to fail. It was also the
881: reason for not including the tantalum TM-measurements in the
882: present analysis. In the eight S2304 series accepted for the
883: determination of the gravitational constant, approximately 8\% of
884: the expected zero-point values could not be determined due to at
885: least one of the three weighings at each load value being rejected
886: by the test for valid weighings.
887:
888:
889: \subsection{Calibration Measurements}
890: \label{sec39} A coarse calibration of the balance was made
891: periodically during the gravitational measurement (before each
892: S288 series) using two calibration masses each with a weight of
893: approximately 100~mg. A correction to the coarse calibration
894: constant due to the nonlinearity of the balance will be discussed
895: in Sec.~\ref{sec310}. The two CM's used for the coarse calibration
896: were short sections of stainless steel wire. The diameter of CM1
897: was 0.50(1)~mm and that of CM2 was 0.96(1)~mm. The surface area of
898: CM1 was approximately 1~cm$^2$ and that of CM2 was 0.5~cm$^2$. The
899: CM's were electrochemically etched to the desired mass and then
900: cleaned in an ethanol ultrasonic bath. The mass of each CM was
901: determined at METAS (Metrology and Accreditation Switzerland) in
902: air with an absolute accuracy of 0.4~\mug\ or a relative
903: uncertainty of 4~ppm. The absolute determinations of the CM masses
904: were made before and after the gravitational measurement with
905: copper TM's and after the second measurement with tantalum TM's.
906: Only the first measurement was used to evaluate the coarse
907: calibration constant employed in the measurement with copper TM's.
908: As will be discussed below, the second and third measurement were
909: used for the measurement with copper TM's only to check the
910: stability of the CM's.
911:
912: A calibration measurement involved either TM and one of following
913: seven additional loads: (1) CM1 alone, (2) CM2 alone, (3) again
914: CM1 alone, (4) empty balance, (5) CM1+CM2, (6) empty balance and
915: (7) CM1+CM2 and nine so called dummy weighings. These measurements
916: were made with no AM's on the balance. After the seventh weighing,
917: a series of nine dummy weighings alternating between upper and
918: lower TM's were made with the AM load set to the value for the
919: next TM weighing. The dummy weighings were made in order to allow
920: the balance to recover from the large load variations experienced
921: during the calibration measurement and thereby come to an
922: approximate equilibrium value before the next TM weighing.
923: Calibrations were made alternately with the upper and lower TM's
924: as load. Calibration measurements were made about twice a day.
925: Including the dummy weighings, each calibration required about 50
926: min.
927:
928:
929: A three-parameter least squares fit was made to the calibration
930: weighings labelled 4,5,6 and 7 above. The fit thereby determined
931: best values for the balance ZP, the slope of the ZP and a
932: parameter representing the effective ZP corrected reading of the
933: balance for the load CM1+CM2. This third parameter is of
934: particular interest since the coarse calibration constant is
935: determined from the known mass of CM1+CM2 (measured by METAS)
936: divided by this parameter. Therefore, the results of the
937: least-squares fit to each set of calibration data gave a value for
938: the coarse calibration constant which then was used to convert the
939: balance output of the S288 series to approximate mass values. An
940: ABA analysis of the first three weighings of each set of
941: calibration data was also made in order to determine the
942: difference in mass between CM1 and CM2.
943:
944:
945: The absolute masses obtained for CM1 and CM2 as determined by
946: METAS are given in columns 2 and 3 of Table~\ref{tab:cm}. Also
947: shown in Table~\ref{tab:cm} (column 4) are the mass difference
948: between CM1 and CM2 as obtained from the METAS measurement in air
949: and the average of our CM measurement in vacuum. The mass
950: differences between CM1 and CM2 measured in vacuum are
951: particularly useful in checking for any mass variation of the
952: CM's.
953:
954: %-------------table 1 CM masses---------------
955: \begin{table}[h]
956: \caption{The mass of the CM's as measured by METAS and the CM1-CM2
957: mass differences measured in air at METAS and in vacuum during the
958: gravitational measurements at PSI. All values are given in $\mu
959: g$.}
960: \label{tab:cm}
961: \begin{ruledtabular}
962: \begin{tabular}{l l l l }
963: \multicolumn{1}{c}{Date} & \multicolumn{1}{c}{CM1} &
964: \multicolumn{1}{c}{CM2} &
965: \multicolumn{1}{c}{Difference} \\
966: \hline
967: Feb 6, 01 & 100,270.30(40) & 100,263.90(40) & 6.40(60)\\
968: Jul. - Sep., 01 & \multicolumn{2}{c}{in vacuum} & {5.853(19)}\\
969:
970: Nov. 29, 01 & 100,270.20(35) &100,262.90(35) & 7.30(50)\\
971: {Jan. - Mar., 02} & \multicolumn{2}{c}{in vacuum} & {7.269(29)}\\
972:
973: {Apr. - May, 02} & \multicolumn{2}{c}{in vacuum} & {7.496(25)}\\
974: May 27, 02 & 100,270.01(35)&100,262.97(35) &7.04(50)\\
975: \end{tabular}
976: \end{ruledtabular}
977: \end{table}
978:
979:
980:
981: It is seen that CM2 mass decreased by 1.00(53)~\mug\ between the
982: first and second METAS measurements while the mass of CM1 was
983: essentially the same in all three measurements. From the mass
984: difference values in air and vacuum it is clear that the change
985: occurred after the measurements with copper TM's ended in
986: Sept.~2001 and before the weighing at METAS in Nov.~2001 which
987: preceded the start of the tantalum measurements. We ascribe this
988: change of CM2 to either the loss of a dust particle or perhaps a
989: piece of the wire itself. The loss of a piece of the wire was
990: possible since the wire used for the CM's had been cut with a wire
991: cutter and there could have been a small broken piece that was not
992: bound tightly to the wire. For this reason only the values given
993: for the first weighing of the CM's were used to determine the
994: coarse calibration constant used for the measurement with copper
995: TM's.
996:
997: A plot of the relative change of the effective ZP corrected
998: balance reading corresponding to the load CM1+CM2 is shown in
999: Fig.~\ref{fig:cm}. It is seen that it changed by only a few ppm
1000: over the 43 days of the measurement. A linear fit made to these
1001: data results in a slope equal to -0.044(6)~ppm/day which is
1002: equivalent to a mass rate variation of -0.0088(12)~\mug\;cm$^{-2}$
1003: \;d$^{-1}$. The uncertainty was obtained by normalizing $\chi^2$
1004: of the fit to the degrees of freedom (DOF).
1005:
1006:
1007:
1008: %--------------fig 11 balance for CM's------------------
1009: \begin{figure}[b]
1010: \includegraphics[width=8.5cm]{cal1_2.ps}
1011: \caption{The change of the effective balance reading for the load
1012: CM1+CM2 as a function of time relative to its value on the first
1013: day. No valid measurements were made between day 229 and 235.}
1014: \label{fig:cm}
1015: \end{figure}
1016:
1017: The slow variation of the effective balance reading for the load
1018: CM1+CM2 seen in Fig.~\ref{fig:cm} could be due either to a change
1019: of the balance sensitivity, to a decrease in the mass of CM1+CM2
1020: due to the removal of a contaminant layer from the CM's in vacuum
1021: or to a combination of both causes. A variation of the balance
1022: sensitivity would have essentially no effect on the analysis of
1023: the weighing for the gravitational measurement as the coarse
1024: calibration constant used for the analysis was determined from the
1025: balance parameter for each S288 series. However, a variation of
1026: the mass of CM1+CM2 would result in an error in the analysis since
1027: the mass would not be the value measured by METAS shown in
1028: Table~\ref{tab:cm}.
1029:
1030: In order to investigate this problem, we have examined the
1031: difference between the balance readings for CM1 and CM2. This
1032: difference is proportional to the surface areas of CM1 and CM2
1033: which differ by approximately a factor of 2 (CM1 area=1~cm$^2$ and
1034: CM2 area=0.5~cm$^2$). The balance reading difference is only
1035: slightly dependent upon the coarse calibration constant so that it
1036: represents essentially the mass difference itself. In
1037: Fig~\ref{fig:cmdiff} is shown the measured mass difference as a
1038: function of time during the gravitational measurement. Also shown
1039: is a linear function fit to these data. The slope parameter of the
1040: fit results in a rate of increase per area equal to
1041: 0.0021(18)~\mug\;cm$^{-2}$ \;d$^{-1}$. The uncertainty has been
1042: determined by normalizing $\chi^2$ to the DOF. The sign of the
1043: slope is such that the CM with the larger area has the larger rate
1044: of increase. A mass difference variation (CM1-CM2) would require a
1045: slope of -0.0088(12)~\mug\;cm$^{-2}$ \;d$^{-1}$. The measured
1046: slope of the effective balance reading for the load CM1+CM2
1047: clearly excludes such a large negative slope as assumed for a mass
1048: variation. We therefore conclude that the variation of this
1049: parameter is due primarily to the sensitivity variation of the
1050: balance.
1051:
1052:
1053: We note that Schwartz \cite{Sch94b} has also found a mass increase
1054: for stainless steel samples in a vacuum system involving a rotary
1055: pump, a turbomolecular pump and a liquid nitrogen cold trap. His
1056: samples were 1~kg masses with surface areas differing by a factor
1057: of 1.8. He measured the thickness of a contaminant layer using
1058: ellipsometry as well as the increase in weight of the sample
1059: during pumping periods of 1.2~d and 0.36~d. The rate of mass
1060: increase per area which he reports is approximately a factor of 5
1061: larger than the value we find. No explanation for this difference
1062: can be made without a detailed knowledge of the partial pressures
1063: of the various contaminant gases in the two systems and the
1064: surface properties of the samples employed.
1065:
1066:
1067:
1068:
1069: %--------------fig 12 CM_diff for copper------------------
1070: \begin{figure}
1071: \includegraphics[width=8.5cm]{cm1cm2.ps}
1072: \caption{The mass difference of the CM's as a function of time
1073: and the linear fit function.} \label{fig:cmdiff}
1074: \end{figure}
1075:
1076:
1077:
1078: There still remains the possibility that a rapid removal of an
1079: adsorbed layer such as water might have occurred between the
1080: absolute determination of the CM masses in air at METAS and the
1081: gravitational measurement in vacuum (i.e. during the pump down of
1082: the system). Schwartz \cite{Sch94a} has measured the mass
1083: variation per unit area of 1~kg stainless steel objects in air
1084: with relative humidity between 3\% and 77\%. He \cite{Sch94b} also
1085: has measured the additional mass variation per area due to
1086: pumping the system from atmospheric pressure at 3\% relative
1087: humidity down to $5\times 10^{-3}$~Pa. His samples were first
1088: cleaned by wiping them with a linen cloth soaked in ethanol and
1089: diethylether and then ultrasonic cleaning in ethanol. After
1090: cleaning, they were dried in a vacuum oven at 50~$^\circ$C. For
1091: these cleaned samples, the weight change found for 3\% to 50\%
1092: humidity variation was 11.5~ng \;cm$^{-2}$ with an additional
1093: change of 29~ng \;cm$^{-2}$ in going from 3\% relative humidity in
1094: air to vacuum (total change of 40.5~ng cm$^{-2}$). Similar
1095: measurements with "uncleaned" samples gave a total change of
1096: 80~ng\;cm$^{-2}$. The variation due to the cleanliness of the
1097: samples was much larger than the difference found for the two
1098: types of stainless steel investigated and the effect of improving
1099: the surface smoothness (average peak-to-valley height equal to
1100: 0.1~\mum\ and 0.024~\mum). Since the cleaning procedure used for
1101: our CM's and their smoothness were different than the samples used
1102: by Schwartz, we have employed the average of Schwartz's "cleaned"
1103: and "uncleaned" objects for estimating the mass change of our
1104: CM's. Based on these data, the relative mass difference found for
1105: both CM's together as measured in air having 50\% humidity and in
1106: vacuum was 0.5~ppm. We assign a relative systematic uncertainty of
1107: this correction equal to the correction itself.
1108:
1109:
1110:
1111:
1112:
1113:
1114:
1115:
1116:
1117:
1118: %------------correction for nonlinearity(might be better as an appendix)------------
1119: \subsection{Nonlinearity Correction}
1120: \label{sec310}
1121: By nonlinearity of the balance, one is referring to the variation
1122: of the balance response function with load, that is, the degree to
1123: which the balance output is not a linear function of the load. The
1124: nonlinearity of a mass comparator similar to the one employed in
1125: the present work has been investigated \cite{Re00} by the firm
1126: Mettler-Toledo. It was found that besides nonlinearity effects in
1127: 10~g load intervals, there was also a fine structure of the
1128: nonlinearity in the 0.1~mg load interval which would be important
1129: for the accuracy of the present measurement. It is the
1130: nonlinearity of our mass comparator in the particular load
1131: interval less than 0.2~g involved in the present experiment that
1132: we wish to determine.
1133:
1134: One expects the nonlinearity of the balance used in this
1135: experiment to be small; however, it should be realized that a 200
1136: mg test mass (two 100~mg CM's) required for having an accurately
1137: known test mass for calibration purposes is over 250 times the
1138: size of the gravitational signal that one wishes to determine. In
1139: addition, the statistical accuracy of the measured gravitational
1140: signal is some 30 times better than the specified accuracy
1141: (2~\mug) of the unmodified commercial balance. One therefore has
1142: no reason to expect the nonlinearity of the balance to be
1143: negligible with this precision. In Sec.~\ref{sec31} we have
1144: presented the general idea that the measurements with 256 AM
1145: values tends to average out the effect of any nonlinearity. We
1146: wish now to give a more detailed analysis of this problem.
1147:
1148: %-------------fig 13 confidence level as function of load------------
1149: \begin{figure}
1150: \includegraphics[width=8.5cm]{confidence_cu.ps}
1151: \caption{The $\chi^2$ probability as a function of the number of
1152: parameters.} \label{fig:chivsl}
1153: \end{figure}
1154:
1155: The correction for nonlinearity makes use of an arbitrary response
1156: as a function of the load. Since the two TM's are essentially
1157: equal ($<400$~\mug\ difference), the variation of the response
1158: function can be thought of as being a function of the additional
1159: load due to the AM's. Although a power series or any polynomial
1160: series would suffice for this function, we have for convenience
1161: used a series of Legendre polynomials
1162: \begin{displaymath}
1163: f(u)=\sum_{\ell=0}^{Lmax} a_\ell P_{\ell} (2 u/u_{max}-1).
1164: \end{displaymath}
1165: The coefficients of $P_\ell$ are chosen subject to the two
1166: conditions that (1) $f(u)=0$ for no load and (2) $f(u)=C$ for
1167: $u=C$ where $C$ is the weight of the two CM's together. These two
1168: conditions represent the sensitivity of the balance over the 0.2~g
1169: range of the calibration (i.e. the coarse calibration). The value
1170: of the maximum load $u_{max}$ in the present measurements was very
1171: nearly $C$. Substituting the above conditions into the response
1172: function, one obtains for the lowest two coefficients the
1173: expressions
1174: \begin{displaymath}
1175: a_0=C/2-\sum_{even \; \ell=2}^{Lmax} a_\ell
1176: \end{displaymath}
1177: and
1178: \begin{displaymath}
1179: a_1=C/2-\sum_{odd \; \ell=3}^{Lmax} a_\ell.
1180: \end{displaymath}
1181:
1182: One can then minimize
1183: \begin{displaymath}
1184: \chi^2=\sum_{n=1}^N \left [f(u_n+s)
1185: -f(u_n)-y_n\right]^2 \sigma_n^{-2}
1186: \end{displaymath}
1187: and thereby determine best values for the parameters $s$ and
1188: $a_\ell$ for $\ell=1$ to $Lmax$. The $y_n$ are the measured
1189: balance signal for the load values $u_n$, $s$ is the load
1190: independent signal and $N$ is the number of different loads with
1191: valid measurements. The error $\sigma_n$ for the load value $u_n$
1192: is the load-independent intrinsic noise of the balance $\sigma_0$
1193: for a single weighing divided by the square root of the number of
1194: weighings for the load $u_n$. The value of $Lmax$ must be chosen
1195: large enough to describe the response function accurately. All of
1196: the parameters in the fit are linear parameters with the exception
1197: of $s$. Thus, there is no difficulty in extending the fit to a
1198: large number of parameters since only the nonlinear parameter must
1199: be determined by a search method.
1200:
1201:
1202:
1203: In order to determine $Lmax$, we calculate the $\chi^2$
1204: probability \cite{Ba89} (often referred to as confidence level) as
1205: a function of $Lmax$. This requires an approximate value for the
1206: intrinsic noise of the balance $\sigma_0$. The value of $\sigma_0$
1207: sets the scale of the $\chi^2$ probability but does not change the
1208: general shape of the function. One can obtain a reasonable
1209: approximation for $\sigma_0$ by setting $\chi^2$ equal to the DOF
1210: obtained for a large number of parameter. We have arbitrarily set
1211: $\chi^2$ equal to the DOF for 61 parameters. The $\chi^2$
1212: probability as a function of the maximum number of parameters is
1213: shown in Fig.~\ref{fig:chivsl}. It is seen that the $\chi^2$
1214: probability reaches a plateau near this maximum number of
1215: parameters.
1216:
1217:
1218: Starting from a low value of $10^{-4}$ for one parameter, the
1219: $\chi^2$ probability rises rapidly to a value of 0.05 for three
1220: parameters. It remains approximately constant at this value up to
1221: 57 parameters where it rises sharply to reach a plateau of
1222: approximately 0.5 at 60 parameters and above. The fit parameter
1223: representing the signal corrected for nonlinearity of the balance
1224: was essentially constant over the entire range of parameters with
1225: a variation of less than $\pm 1.3$~ng. The signal for one
1226: parameter representing complete linearity was 784.8994~\mug. The
1227: signal of the plateau region from 60 to 67 parameters was
1228: 784.9005~\mug\ with a statistical uncertainty of 5.5~ng. In this
1229: region the signal varied by less than 0.2~ng. We therefore take
1230: the nonlinearity correction of the measured signal to be
1231: 1.1(5.5)~ng (i.e. the difference between the signal using one
1232: parameter as would be obtained with no correction and the average
1233: value obtained for 60 to 67 parameters).
1234:
1235:
1236: %----------------fig 14 linearity data and fit-------------
1237: \begin{figure}
1238: \includegraphics[width=8.5cm]{response_residual.ps}
1239: \caption{Signal and fit function employing 60 parameters as a
1240: function of load. The data are shown as a stepped line. The fit is
1241: the smooth curve. The lower plot shows the normalized residuals.
1242: Residuals were divided by the relative uncertainty of each point.
1243: The normalization has been chosen such that the rms value of the
1244: residuals is 1.} \label{fig:sigfit}
1245: \end{figure}
1246:
1247: The nonlinear signal and fit as a function of load determined for
1248: 60 parameter is shown in Fig.~\ref{fig:sigfit}. The function shows
1249: many narrow peaks with widths of 3 to 10 load steps and with
1250: amplitudes of roughly 0.1~\mug. In principle one could use this
1251: response function to correct the individual weighings with various
1252: loads; however, we prefer to use the signal as corrected for
1253: nonlinearity over the entire range of measurements as described
1254: above. The variation of the response function indicates that a
1255: measurement made at an arbitrary load value could be in error by
1256: as much as $\pm 130$~ng assuming the response to be linear. This
1257: is to be compared with the assumed uncertainty in ref.~\cite{Sb98}
1258: due to nonlinearity of $\pm 200$~ng.
1259:
1260:
1261:
1262:
1263:
1264:
1265: %-------------Section Correction of TM Sorption Effect----------
1266: \subsection{Correction of the TM-Sorption Effect}
1267: \label{sec311}
1268: Moving the FM's changed slightly the temperature of the vacuum
1269: tube surrounding the TM's. These temperature variations were due
1270: to changes in the air circulation in the region of the vacuum tube
1271: as obstructed by the FM's. An increase of the wall temperature of
1272: the tube caused adsorbed gases to be released which were then
1273: condensed onto the TM. Since the temperature variation was
1274: different in the regions near the upper and lower TM's, this
1275: resulted in a variation of the weight difference between the upper
1276: and lower TM's (i.e. a "false" gravitational signal).
1277:
1278: %-----------fig 15 temperature at position of TM's----------
1279: \begin{figure}
1280: \includegraphics[width=8.5cm]{temps.ps}
1281: \caption{Temperatures of the vacuum tube measured at the position
1282: of the TM's. The upper curve is the temperature at the position of
1283: the upper TM. The square wave in the middle section of the plot
1284: indicates the FM motion. The data (crosses) for the lower TM and
1285: fit function (solid line) are shown in the lowest section of the
1286: figure.} \label{fig:temp}
1287: \end{figure}
1288:
1289:
1290: The temperature variation at the positions of the upper and lower
1291: TM's during one day of the gravitational measurement is shown in
1292: Fig.~\ref{fig:temp} along with a curve representing the FM motion.
1293: The peak-to-peak temperature variation was approximately
1294: $0.04^\circ \mbox{C}$ at the upper position and $0.01^\circ
1295: \mbox{C}$ at the lower position. The shape of the temperature
1296: variation at the upper position was used as a fit function
1297: (employing an offset and an amplitude parameter) to obtain a
1298: better determination of the temperature variation at the lower
1299: position. There were 32 one-day measurements of the temperature
1300: variations during the gravitational measurement. The average
1301: amplitude at the lower position determined from these 32
1302: measurements was $0.0138(2)^\circ \mbox{C}$.
1303:
1304: The signal produced by these temperature variations was small and
1305: therefore not directly measurable with the balance in a reasonable
1306: length of time. The procedure that was employed to determine this
1307: temperature dependent signal was to use four electrical heater
1308: bands to produce a variation of the temperature distribution along
1309: the vacuum tube that was a factor of approximately seven larger
1310: than the variation resulting from the motion of the FM's. The
1311: bands were positioned 30 cm above and below the positions of the
1312: upper and lower TM's. The heater windings were bifilar to avoid
1313: magnetic effects. The heater power (less than 3~W total) was
1314: turned off and on with the same 8-hour period as the FM motion and
1315: produced essentially no change in the average temperature of the
1316: vacuum tube in the day-long measurement. The FM's were not moved
1317: during the measurements with heaters. The signal (TM weight
1318: difference as determined with the balance) obtained during a one
1319: day measurement with heaters is shown in Fig.~\ref{fig:sorption}.
1320: The shape of the fit function (employing an offset and an
1321: amplitude parameter) shown in this figure was obtained from the
1322: variation of the temperature difference at the upper and lower
1323: positions of the TM's. The signal obtained from the fitting
1324: procedure was 0.114(40)~\mug.
1325:
1326: %---------------fig 16 sorption fit-----------
1327: \begin{figure}
1328: \includegraphics[width=8.5cm]{signal_run3.ps}
1329: \caption{Weight difference between TM's as a function of time for
1330: a temperature variation roughly 10 times that of the gravitational
1331: measurement. The solid curve is the best fit of the temperature
1332: variation difference at upper and lower TM positions. For the
1333: purpose of this plot, an arbitrary offset of the weight difference
1334: between upper and lower TM has been employed.}
1335: \label{fig:sorption}
1336: \end{figure}
1337:
1338: In order to scale the heater produced signal to that resulting
1339: from the FM motion during the gravitational measurement, we make
1340: the simplifying assumption that the signal variation is
1341: proportional to the temperature variation at the upper TM
1342: position minus the temperature variation at the lower TM
1343: position. The term variation in this statement refers to the
1344: variation about its mean value. One uses the temperature
1345: difference since the signal is defined as the difference between
1346: TM weighings.
1347:
1348: With just four heater bands it was not possible to obtain a
1349: variation of the temperature distribution along the vacuum tube
1350: that was exactly a constant factor times that of the FM motion.
1351: For the best adjustment that we were able to obtain, the ratio of
1352: the heater produced temperature variation to the FM produced
1353: variation was 7.1 at the upper position and 9.2 at the lower. The
1354: ratio for the variation of the temperature difference at the upper
1355: and lower positions relative to the gravitational values was 6.8.
1356: These ratios are based on the peak-to-peak amplitudes obtained for
1357: the fitted functions. The scaling factor for the temperature
1358: difference ratio is the reciprocal of the temperature difference
1359: ratio or 0.147. This results in a scaled signal of 0.0168(58)~\mug\
1360: where the uncertainty is the statistical uncertainties of the
1361: measured signal and the scaling factor. The scaled signal ("false"
1362: signal) is to be subtracted from the total signal measured in the
1363: gravitational experiment.
1364:
1365: In order to check our assumption regarding the scaling factor, we
1366: have made four additional one-day measurements in which the
1367: temperature variations were very different from that produced by
1368: the FM motion. The object of these measurements was to determine
1369: whether the scaled signals obtained with the heaters were
1370: consistent with one another when calculated with the assumed
1371: scaling factors. The most extreme distribution involved a
1372: temperature variation of the lower TM which was even larger
1373: (factor of 4) than that of the upper TM. The signals obtained in
1374: all of the test measurements were consistent with each other
1375: within their statistical uncertainties (relative uncertainties of
1376: approximately 30~\%). We therefore conclude that the assumption
1377: used for scaling the signals was sufficiently accurate for the
1378: present purpose. Nevertheless, we assign a systematic uncertainty
1379: to the scaled signal equal to its statistical uncertainty of
1380: 5.8~ng (relative systematic uncertainty of the "false" signal is
1381: 35~\%).
1382:
1383: %---------section Magnetic Forces----------------
1384: \subsection{Magnetic Forces on the Test Masses}
1385: In the absence of a permanent magnetization, the $z$ component of
1386: force on the TM due to a magnetic field can be calculated from
1387: \begin{displaymath}
1388: F_z=-\mu_0 \chi_m V H \frac{\partial H}{\partial z}
1389: \end{displaymath}
1390: where $V$ denotes the volume of the TM, $\chi_m$ is its magnetic
1391: susceptibility and $H$ is the magnetic field intensity. The
1392: magnetic properties of the TM's were measured by METAS. No
1393: permanent magnetization was found ($<0.08$~A/m). The magnetic
1394: susceptibility was $4\times 10^{-6}$ for the copper TM's. The
1395: magnetic field intensity for both positions of the FM's was
1396: measured at cm intervals along the axis of the vacuum tube at the
1397: positions occupied by the TM's using a flux gate magnetometer. The
1398: difference of $F_z$ for the FM positions obtained from these data
1399: was 0.01~ng which is a negligible correction to the measured
1400: gravitational signal.
1401:
1402: %--------section tilt of balance-------------
1403: \subsection{Tilt Angle of Balance}
1404: Since the weight of the TM's and the weight of the CM's both
1405: produce forces on the balance arm in the vertical direction, a
1406: small angle between the balance weighing direction and the
1407: vertical produces no error in the weighing of the TM's. However,
1408: if the balance weighing direction is correlated with the motion of
1409: the FM's, a systematic error in the measured gravitational signal
1410: will result. Sensitive angle monitors were mounted on the base of
1411: the balance. No angle variation correlated with the motion of the
1412: FM's was found with a sensitivity of 100~nrad. Since the
1413: sensitivity of the balance varies with the cosine of the angle
1414: (near 0 rad), this limit is completely negligible. For a balance
1415: misalignment of 0.01~rad relative to vertical and a correlated
1416: variation of 100~nrad with respect to this angle due to the FM
1417: motion, the relative signal variation is approximately 0.001~ppm.
1418:
1419:
1420: %-------------------Mass Integration------------------------
1421: \section{Determination of the Mass-Integration Constant}
1422:
1423: \label{sec4}
1424: One must relate the gravitational constant to the measured
1425: gravitation signal. This involves integrating an inverse square
1426: force over the mass distribution of the TM's and FM's. The
1427: gravitational force $F_z$ in the z (vertical) direction on a
1428: single TM produced by both FM's is given by
1429: \begin{equation}
1430: F_z=G \int\int \frac {\bf{e_z} \cdot(\bf{r_2-r_1}) \; \it{dm_1
1431: dm_2}}{|\bf{r_2-r_1}|^{\it3}}
1432: \end{equation}
1433: where $\bf{e_z}$ is a unit vector in the $z$ direction, $\bf{r_1}$
1434: and $\bf{r_2}$ are vectors from the origin to the mass elements
1435: $dm_1$ of the TM and $dm_2$ of the FM's and $G$ is the
1436: gravitational constant to be determined. The mass-integration
1437: constant is the double integral in Eq. (1) multiplying $G$.
1438: Actually, the mass-integration constant for the present experiment
1439: is composed of four different mass-integration constants, namely
1440: those for the upper TM and lower TM with the FM's together and
1441: apart. We shall use as mass-integration constant the actual
1442: constant multiplied by the 1986 CODATA value of $G$
1443: ($6.67259~\Gunit\) and give the result in dimensions of grams
1444: "force" (i.e. the same dimensions as used for the weighings).
1445:
1446: The objects contributing most to $F_z$ (TM's, FM tanks and the
1447: mercury) have very nearly axial symmetry which greatly simplifies
1448: the integration. Parts which do not have axial symmetry were
1449: represented by single point masses for small parts and multiple
1450: point masses for larger parts. For axial symmetric objects, we
1451: employ the standard method of electrostatics for determining the
1452: off-axis potential in terms of the potential and its derivatives
1453: on axis (see e.g.\cite{Gl56}). The force on a cylindrical TM in
1454: the $z$ direction produced by an axially symmetric FM can be
1455: conveniently expressed as (see Eq.~\ref{eq:massint_main},
1456: Sec.~\ref{sec7})
1457:
1458: \begin{align}
1459: &F_z = 2 M_{TM} \times \notag
1460: \\
1461: &\sum_{n=0}^\infty V_0^{(2n+1)}
1462: \sum_{i=0}^n \frac{1}{\left( -4 \right)^i}
1463: \frac{1}{i!\left(i+1\right)!} \frac{1}{\left( 2 n- 2 i +1
1464: \right)!} b^{ 2 n -2 i } r^{2i} \
1465: \end{align}
1466: where $M_{TM}$ is the mass of a cylindrical TM with radius r and
1467: height b, and $V_0^{(2n+1)}$ is the $2n+1$st derivative of the
1468: gravitational potential with respect to $z$ evaluated at the
1469: center of mass of the TM ($r=0$, $z=z_0$).
1470:
1471: The potential $V(r=0,z)$ of the various FM components having axial
1472: symmetry was determined analytically for three types of axially
1473: symmetric bodies, namely a hollow ring with rectangular cross
1474: section, one with triangular cross section and one with circular
1475: cross section. This allows one to calculate the gravitation
1476: potential of the tank walls and the mercury content of the tank as
1477: a sum of such bodies. For example, the region between measured
1478: heights on the top plate and $z=0$ at two values of the radius was
1479: represented by a cylindrical shell composed of a right triangular
1480: torus and a rectangular torus (i.e. a linear interpolation between
1481: the points describing the cross section of the rings). O-rings
1482: were calculated employing the equation for rings with circular
1483: cross sections. A total of nearly 1200 objects (point masses and
1484: rings of various shapes) were required to describe the two FM's.
1485:
1486: The derivatives of the potential were evaluated using a numerical
1487: method called ``automatic differentiation" (see e.g. \cite{Ra81}).
1488: For the geometry of the present experiment, the terms in the
1489: summation over $n$ decrease rapidly so that 8 terms were
1490: sufficient for an accuracy of 0.02~ppm in the mass-integration
1491: constant.
1492:
1493:
1494: \subsection{Positions of TM's and FM's}
1495: \label{sec41} In order to carry out the mass integration, one
1496: needs accurate weight and dimension measurements of the TM's and
1497: FM's as well as distances defining their relative positions. The
1498: dimension and weight measurements for TM's were described in
1499: Sec.~\ref{sec32}. The measurement of the TM positions shown in
1500: Fig.~\ref{fig:geom} will now be addressed.
1501:
1502: %-----------------fig 18 vertical positions--------------
1503: \begin{figure}[htb]
1504: \includegraphics[width=8.5 cm]{geom.eps}
1505:
1506: \caption[Geometry.]{Drawing showing the measured
1507: vertical distances to TM and FM surfaces for the two FM positions
1508: (T=Together and A=Apart).} \label{fig:geom}
1509: \end{figure}
1510:
1511: A special tool was made to adjust the length of the tungsten wires
1512: under tension. Each wire made a single loop around the post on
1513: either side of the TM and a thin tube was crimped onto the wires
1514: to hold them together thereby forming the loop (see
1515: Fig.~\ref{fig:cotm}). The position of the TM could only be
1516: measured with the vacuum tube vented. The vacuum tube was removed
1517: below a flange located at a point just above the upper TM. The TM
1518: hanging from the balance was then viewed through the telescope of
1519: an optical measuring device to determine its position.
1520:
1521: The vertical position of the TM's and FM's was measured relative
1522: to a surveyor's rod which was adjusted to be vertical. The bottom
1523: of the surveyor's rod was positioned to just touch a special
1524: marker mounted on the floor of the pit. The surveyor's rod had
1525: accurate markings every cm along its length. A precision levelling
1526: device in which the optical axis of the telescope could be
1527: displaced by somewhat more than a cm was then used to compare the
1528: position of the upper surface of a TM with a mark on the
1529: surveyor's rod. Similar measurements were made for the FM's. These
1530: measurements were made before and after each of the three
1531: gravitational measurement(Cu, Ta 1 and Ta 2 TM's). Although the
1532: measuring device including the surveyor's rod was removed from the
1533: pit after each of these measurements, the reproducibility of each
1534: position measurement was found to be better than 35 $\mu m$. The
1535: accuracy of the average of the two sets of position measurement
1536: for each type of TM including systematic uncertainty was 35~\mum.
1537:
1538: A small vertical displacement of the TM's occurred when the system
1539: was evacuated. This was measured during the evacuation of the
1540: system by observing the TM's through the windows on the side of
1541: the vacuum tube using the levelling device that was also used for
1542: measuring the TM position in air. The vertical displacement was
1543: measured several times and found to be 0.10(3)~mm. This
1544: displacement is shown as $\varepsilon$ in Fig.~\ref{fig:geom}).
1545:
1546: The angle of the TM axis relative to the vertical direction was
1547: also determined with the same precision levelling instruments by
1548: measuring the height on the top surface of the TM at two opposite
1549: points near the outer radius of the TM. This was done for each TM
1550: from a viewing direction almost perpendicular to the plane of the
1551: supporting wires. The angle of the axis relative to the vertical
1552: was found to be less than $1^\circ$ for both TM's.
1553:
1554: The horizontal positions of the TM's were determined using a
1555: theodolite. The left and right sides of the TM were viewed through
1556: the telescope of the theodolite relative to an arbitrary zero
1557: angle. The horizontal position of the central tube was determined
1558: relative to the same zero angle. These measurements were made for
1559: each TM and FM from two nearly perpendicular viewing angles. The
1560: measurements were made before and after each gravitational
1561: measurement. The radial positions of the upper and lower TM's
1562: relative to the common axis of the FM's were found to be 0.45~mm
1563: and 0.50~mm, respectively. The overall accuracy of the radial
1564: positions of the TM's from these measurements was 0.1~mm. This
1565: uncertainty results in only a small uncertainty in the value of G
1566: determined in this experiment due to the extremum of the force
1567: field in the radial direction. No problem was experience with
1568: pendulum oscillation during these measurements as the amplitudes
1569: were strongly damped in air.
1570:
1571: \subsection{Dimensions and Weight of the FM's}
1572: \label{sec42} The individual parts of the FM's were weighed at
1573: METAS with an accuracy of 1~g. The weights were corrected for
1574: buoyancy to obtain the masses.
1575:
1576: The narrow confines of the pit made measurement of the FM's
1577: deformed by the mercury load difficult. Although measurements of
1578: the individual pieces before assembly were in principle more
1579: accurate, the loading and temperature difference between dimension
1580: measurements and gravitational measurement reduced the accuracy of
1581: these measurements. In addition, it is known that long term
1582: loading can release tensions in the material which result in
1583: inelastic deformation of the material. Therefore, the measurements
1584: made with mercury load were always used in the analysis when
1585: available.
1586:
1587: The uncertainty in the height of the central piece proved to be
1588: very important in determining the uncertainty of the
1589: mass-integration constant. Due to the various types of
1590: measurements for this dimension with different systematic effects,
1591: we decided that the best value would be an equally weighted
1592: average of the four available measurements with the uncertainty
1593: being determined from the rms (root-mean-square) deviation from
1594: the mean. The measurements employed were the following: (1) a
1595: Coordinate Measuring Machine (CMM) measurement before the tanks
1596: were assembled, (2) a Laser Tracker (LT) measurement with mercury
1597: loading during the experiment, (3) a LT measurement in the machine
1598: shop with no loading after completion of the experiment and (4) a
1599: CMM measurement after the tanks had been disassembled at the end
1600: of the experiment. The two CMM measurements were independent in
1601: that they were made with different CMM devices and with different
1602: temperature sensors. The uncertainty in the height as determined
1603: from the rms deviation was 19~\mum\ for the upper tank and 9~\mum\
1604: for the lower tank.
1605:
1606: %----------------fig 19 cut away view of tank-----------------
1607: \begin{figure}[htb]
1608: \includegraphics[width=8.5cm]{tank.eps}
1609: \center{figure} \caption[Drawing of the field mass.]{Drawing of
1610: the field mass. All dimensions are given in mm.} \label{fig:tank}
1611: \end{figure}
1612:
1613: A cut-away drawing of a FM tank is shown in Fig.\ref{fig:tank}. All pieces
1614: were made from stainless steel type No.~1.4301 which is resistant
1615: to mercury and has a low magnetic susceptibility. The pieces were
1616: sealed to one another with mercury resistant Perbunan O-rings. The
1617: top and bottom plates were fastened to the central piece with 12
1618: screws. The top and bottom plates were screwed to the outer casing
1619: with 24 screws.
1620:
1621: Due to the nearness of the central piece to the TM's, especially
1622: close tolerances were specified for this piece. The central piece
1623: was annealed before machining to remove tensions which could
1624: deform the piece during machining. A surface roughness value of
1625: $<1$ $\mu m$ was obtained by grinding the surface after machining.
1626: The roughness value is defined as the average height of the peaks
1627: times the area of peaks relative to the total area of the surface.
1628:
1629: The top and bottom plates were less critical than the central
1630: piece. A surface roughness value of 6 to 10~\mum\ was specified
1631: for the machining of the inner surface of these pieces. The inner
1632: surface of the casing was even less critical and a surface
1633: roughness value of 15~\mum\ was specified for it.
1634:
1635:
1636: Since the mercury was filled into the tanks under vacuum
1637: conditions, it filled more or less the exact surface profile (i.e.
1638: the region between the grooves caused by machining). The mercury
1639: in the filled tank was under positive pressure on all surfaces
1640: including the top plate. The over pressure ranged from about 0.1
1641: bar on the top plates (due to the height of mercury in the
1642: compensation vessel shown in Fig.~\ref{fig:tank}) and 1 bar on the
1643: bottom plates. The bulging of the central cylinder and the outer
1644: casing due to the mercury pressure was calculated using the
1645: equations for thin cylindrical shells \cite{Ti87}. The inward
1646: bulging of the central tube was found to have a maximum value of
1647: 0.17~\mum\ which is completely negligible for the present
1648: experiment. The maximum outward bulging of the outer casing was
1649: approximately 4~\mum\ and resulted in a small (8~ppm) correction
1650: of the volume. The loading of the tanks produced a 2~\mum\
1651: elongation of the outer casing corresponding to a relative volume
1652: increase of 1~ppm.
1653:
1654: Before the tanks were assembled, measurements of the individual
1655: pieces were made with a coordinate measuring machine (CMM). The
1656: inner diameters of the central tube were checked with a special
1657: dial gauge and found to agree with the CMM values. The wall
1658: thickness of the outer casing, top and bottom plates were measured
1659: with a wide jaw micrometer having a dial gauge readout.
1660:
1661: %---------------figs 17 LT measurement----------
1662: \begin{figure}
1663: \includegraphics[width=8.5cm]{de2_real.ps}
1664: \caption{Data and fit for the upper surface of the lower tank as a
1665: function of radius. The solid curve is the fit function based on
1666: the theory of circular plates. The offset has been chosen such
1667: that the z-value is 0~mm at a radius of 60~mm.}
1668: \label{fig:lt}
1669: \end{figure}
1670:
1671: After the tanks were filled with mercury, measurements of the
1672: outer surface of the top and bottom plates were made with a
1673: laser-tracker (LT). It was not possible to measure the outer
1674: surface of the casing due to the close confines of the pit. The LT
1675: measurements were made in the dynamic mode in which the retro
1676: reflector is moved on the surface and readings are taken as fast
1677: as possible (1000 per s). An example of these data for the top
1678: surface of the lower tank is shown in Fig.~\ref{fig:lt}. The data
1679: for the other plates are similar. The force due to the mercury
1680: loading which tended to stretch the central tube, and to a lesser
1681: extent the outer casing, could be determined from these data using
1682: equations based on thin axial symmetric plates and shells
1683: \cite{Ti87}. The force on the central tube was calculated to be
1684: ($17\pm 8$~kN) which resulted in an elongation of 15~\mum.
1685:
1686: Finally, after the mercury had been removed from the tanks,
1687: additional LT measurements were made of the outside height of the
1688: central tube in order to clear up a discrepancy of this dimension
1689: as measured with the CMM and LT.
1690:
1691: The form of the central tube and outer casing as determined by the
1692: analysis of the measurements was very nearly circular; however,
1693: the deviation from perfect roundness was larger than the expected
1694: accuracy of the measurements. For the purpose of determining the
1695: actual accuracy of the measurements, a least-squares fit was made
1696: to the data employing a number of Fourier terms. The uncertainty
1697: of the measured data was then determined by setting the $\chi^2$
1698: of this fit equal to the DOF.
1699:
1700: For the purpose of determining the uncertainty of the volume and
1701: the mass-integration constant, we have employed effective
1702: dimensions describing a hollow, right circular cylinder. The
1703: effective value of the small radius $r$ and height $h$ of the
1704: central piece were assigned the approximate values of 60 mm and
1705: 650 mm, respectively. The effective inner radius $R$ of the casing
1706: was then determined such that the inside volume of the tank was
1707: the value determined from measurements.
1708:
1709: Besides the accuracy of the directly measured dimensions, one has
1710: also to consider the effect of the expansion coefficient of the
1711: stainless steel times the temperature difference between
1712: measurement of the piece and the temperature during the
1713: gravitational measurement, accuracy of the surface roughness and
1714: the deformation due to the load. The accuracy of the temperature,
1715: roughness and deformations effects are assumed to be one half of
1716: the change caused by these effects. An estimate of these
1717: accuracies for the inside effective dimensions of the tanks is
1718: given in Table~\ref{tab:tankaccuracy}.
1719:
1720: %------------Table 2 accuracy of inside tank dimensions-----------
1721: \begin{table}[tbp]
1722: \caption{Estimated uncertainty in $\mu m$ of the effective values
1723: for the radius $r$ of the central tube, the radius $R$ of the
1724: outer casing and the height $h$ of the central tube due to various
1725: effects. The uncertainties apply to the inside dimensions of the
1726: tank.}
1727: \label{tab:tankaccuracy}
1728: \begin{ruledtabular}
1729: \begin{tabular}{l r r r}
1730: \multicolumn{1}{l}{Description}&
1731: \multicolumn{1}{c}{$\sigma_r$} &
1732: \multicolumn{1}{c}{$\sigma_R$} & \multicolumn{1}{c}{$\sigma_h$}\\
1733: \hline measurement & 1.0 & 5 & 20
1734: \\thickness & - & - & 5
1735: \\temperature & 0.3 & 4 & 6
1736: \\elongation & - & - & 7
1737: \\bulge & 0 & 2 & -
1738: \\roughness & 0 & 6 & 7
1739: \\ \hline
1740: Total & 1.1 & 9 & 24
1741: \\
1742: \end{tabular}
1743: \end{ruledtabular}
1744: \end{table}
1745:
1746:
1747:
1748: \subsection{Weighing the Mercury}
1749: \label{sec43}
1750: After a preliminary measurement in which the tanks were filled
1751: with water, the water was drained from the tanks and they were
1752: ventilated with warm dry air. The tanks were then evacuated to a
1753: pressure of $10^{-2}$~mbar using a rotary pump with an oil filter
1754: to prevent back streaming in preparation for being filled with
1755: mercury.
1756:
1757: Since the tanks were to be filled with mercury only once, every
1758: effort was made to weigh the mercury as accurately as possible
1759: during the filling. The mercury (specified purity 99.99\%) was
1760: purchased in 395 flasks each weighing 36.5~kg and containing
1761: 34.5~kg of mercury. The general procedure for filling each tank was as
1762: follows. The outer surface of the flasks were first cleaned with
1763: ethanol. Half of the flasks were brought into a measuring room
1764: near the experiment and allowed to come into equilibrium with the
1765: room atmosphere for a few days. These flasks were then weighed
1766: over a period of one week. Then, one after another, the flasks
1767: were attached to the transfer device. Most of the mercury in a
1768: flask was transferred via compressed nitrogen, first into an
1769: intermediate vessel used as a vacuum lock and then into the
1770: evacuated tank. A small amount of mercury was intentionally left
1771: in each flask in order not to transfer any of the thin oxide layer
1772: floating on the surface. The filling of a tank required one week.
1773: After completion of the filling, the flasks with their small
1774: remaining mercury content were weighed again. The entire process
1775: was then repeated for the second tank.
1776:
1777: Various precautions and test measurements were undertaken to
1778: insure the accuracy of the weighings. The balance employed for
1779: these measurements was type SR 30002 made by Mettler-Toledo. The
1780: balance was operated in the differential mode with accurately
1781: known (5~mg) standards for weights less than 2~kg (for empty
1782: flasks) and a stainless steel mass of approximately 36.6~kg made
1783: in our machine shop (for full flasks). The mass of the 36.6~kg
1784: weight was calibrated several times at METAS and remained constant
1785: within the 18~mg (the certified accuracy of the weighings) during
1786: the weighing of the mercury. The reproducibility of the weighing
1787: of an almost empty flask or full flask was 20~mg. The average of
1788: three weighings was made for empty and full flasks with the
1789: balance output transferred directly to a computer via an RS232
1790: interface. The balance was checked for nonlinearity and none was
1791: found within the accuracy of the standard weights. A centering
1792: table was used which allowed the flasks to be off center by as
1793: much as 2~cm without influencing the measurement. Atmospheric data
1794: used for buoyancy corrections were taken several times a day. A
1795: 10~kg calibration test was made once a day and the balance zero
1796: was checked every hour. It was found that the flasks were
1797: magnetized along their symmetry axis. Rotating the flask on the
1798: balance did not change the measured weight but inverting the flask
1799: resulted in a 100~mg difference. Since weighings were always made
1800: with the flasks in an upright position, no magnetization error
1801: occurred in the difference between full and empty flasks. The
1802: variation of the weight for 12 flasks was monitored over a period
1803: of 8 weeks. The variation was similar for all 12 flasks and
1804: amounted to about 20~mg per flask during the 3 weeks required to
1805: weigh the flasks and fill the tank. Mercury droplets which had not
1806: reached the tanks and small flakes of paint which had accidentally
1807: been removed from the outer surface of the flasks during the
1808: transfer process were collected and weighed.
1809:
1810:
1811: The total uncertainty in the approximately 6760~kg of mercury in
1812: each tank was 12~g for the upper tank and 15~g for the lower tank
1813: which gives a relative accuracy of 2.2~ppm for the mass of mercury
1814: in the tanks. A listing of the various weighing uncertainties is
1815: given in Table 3. An estimate of the mercury and flask residue
1816: that had not been collected (and the uncertainty of the amount
1817: collected) was assumed to be 20\% of the amount collected. The
1818: pressurizing of the flasks with nitrogen during transfer resulted
1819: in a buoyancy correction of the almost empty flasks due to the
1820: difference in density between air and nitrogen. The flasks were
1821: sealed after use, but air gradually replaced the nitrogen. The
1822: assumption of a constant density (approximate equation) for the
1823: gas left in each empty flask during the time of the measurement
1824: amounted to an uncertainty of 7 and 8~g for the upper and lower
1825: tanks, respectively. The change in mass of the flasks during the
1826: three week measurements gave an uncertainty of 4~g for each tank.
1827: The accuracy of the standard masses caused an uncertainty of 3.7~g
1828: for each tank. Air buoyancy uncertainties resulted in an
1829: uncertainty of 1.2~g for each tank. The statistical uncertainty
1830: due to reproducibility of the balance was 0.4~g for each tank.
1831: %----------table 3 weighing uncertainties---------------
1832: \begin{table}[tbp]
1833: \caption{A listing of the weighing uncertainties for the upper and
1834: lower tanks. All uncertainties except that of the standard masses
1835: are independent for each tank. The total mass of mercury in each
1836: tank was approximately 6760 kg.}
1837: \label{tab:mercury}
1838: \begin{ruledtabular}
1839: \begin{tabular}{ l c c }
1840: \multicolumn{1}{ l }{Type of Uncertainty}&
1841: \multicolumn{1}{c}{Upper}& \multicolumn{1}{c}{Lower}\\
1842: &
1843: \multicolumn{1}{c}{Tank[g]}& \multicolumn{1}{c}{Tank[g]}\\
1844:
1845:
1846: \hline loss of mercury and residue from flasks & 8 & 11
1847: \\approximate equation & 7 & 8
1848: \\mass variation of flasks & 4 & 4
1849: \\uncertainty of standard masses & 3.7 & 3.7
1850: \\buoyancy correction & 1.2 & 1.2
1851: \\ weighing statistics & 0.4 & 0.4
1852: \\ \hline total uncertainty & 12 & 15
1853: \end{tabular}
1854: \end{ruledtabular}
1855: \end{table}
1856:
1857: \subsection{Mass Distribution of the Central Piece}
1858: \label{sec44}
1859: Although the principle mass making up the FM's was mercury and
1860: therefore had only a negligible density variation due to its
1861: pressure gradient, the density variation of the walls of the tanks
1862: was important in determining the mass-integration constant. In
1863: particular, the two central pieces which were very near to the
1864: TM's and which were composed of three different pieces of
1865: stainless steel welded together were critical for this
1866: determination. Therefore after the gravitational measurements had
1867: been completed, the central pieces were cut into a number of rings
1868: in order to determine the density of these rings and thus the mass
1869: distribution of the central pieces.
1870:
1871: As shown in Fig.~\ref{fig:tank}, each central piece was composed
1872: of upper and lower flanges and a central tube. The material of the
1873: flanges extended about 40~mm beyond the surface of the flanges in
1874: the form of the central tube. In order to determine the vertical
1875: mass distribution in the flanges, three rings of 10~mm height were
1876: cut from each of these 40~mm sections with the last ring
1877: straddling the weld between flange and central tube. The weight
1878: and dimensions of the various pieces (12 rings, 4 flanges and two
1879: central sections of the tube) were used to determine the densities
1880: of these pieces. The densities of the flanges were found to be
1881: between 7.9138 to 7.9147~g cm$^{-3}$ and the densities of the
1882: central section of tubes were 7.9062 and 7.9101~g cm$^{-3}$. The
1883: accuracy of the absolute-density determinations was somewhat
1884: better than 0.001~g cm$^{-3}$. The density of the weld regions did
1885: not differ significantly from that of the flanges. The variation
1886: of the flange densities over the 65~mm of the flange and adjoining
1887: section of the tube was found to be less than 0.005~g cm$^{-3}$.
1888: From these measurements, it was not possible to determine a radial
1889: density gradient of the flanges. For calculating the effect of a
1890: radial gradient on the mass-integration constant, we shall make
1891: the assumption that the radial density variation was less than
1892: 0.005~g cm$^{-3}$ over the radial dimension of the flange
1893: (160~mm). The vertical density gradients of the central tubes were
1894: not important since their effects on the gravitational signal are
1895: almost completely cancelled due to the symmetry of central tube
1896: relative to the apart-together measurements with the FM's.
1897:
1898:
1899:
1900:
1901:
1902:
1903:
1904: \subsection{Using the Measured Dimensions}
1905: \label{sec45}
1906: The first step in calculating the mass-integration constant was to
1907: enter the measured dimensions and masses of the various pieces in
1908: a computer program. For pieces that had essentially axial symmetry
1909: such as the central tube and the outer casing, an average radius
1910: was determined from the data measured at each height and used in
1911: further calculations. For horizontal surfaces which were nearly
1912: planar such as the top and bottom plates, an average height was
1913: determined from the data at each measured radius and used in
1914: further calculations. Since the original data were normally
1915: available only in Cartesian coordinates, it was necessary to
1916: determine the symmetry axis and make the conversion to cylindrical
1917: coordinates. For data without axial symmetry such as screws, screw
1918: holes and linear objects, single or multiple point masses were
1919: used. With this reduced set of dimensions, approximately 580 data
1920: elements were necessary to describe each tank.
1921:
1922: As a preliminary calculation related to the mass-integration
1923: constant, the volume of the individual pieces and the inside
1924: volume of the tanks were calculated from the reduced dimensions.
1925: Using the known weight of the piece, the calculated volume allowed
1926: the density of the material to be determined. This was a valuable
1927: test to check whether the input data for the piece was reasonable.
1928:
1929: The volume for pieces with axial symmetry was determined by making
1930: a linear interpolation between the points of the reduced data. The
1931: volume of a piece was thus composed of a sum of cylindrical rings
1932: with rectangular and right triangular form. A cylindrical ring
1933: with circular cross section was used for the volume of O-rings.
1934: The accuracy of the linear approximation in the volume
1935: determination was checked relative to a quadratic approximation of
1936: the surface. The linear approximation was found to be sufficient
1937: for all calculations.
1938:
1939:
1940: The original CMM measurements had been made for 12 $\varphi$
1941: angles at 14 heights on the central tube, at 4 radii on the
1942: horizontal surfaces of the central flange, at 11 radii on the
1943: horizontal surfaces of the top and bottom plates and at 7 heights
1944: on the outer casing. Although many more points of the horizontal
1945: surfaces were measured with the LT, they were reduced to the
1946: original CMM points for the purpose of volume integration and mass
1947: integration by fitting functions to the LT data. Only outside
1948: surfaces were measured with the LT. Inside dimensions were
1949: obtained from the LT data by subtracting the micrometer-thickness
1950: values. The only measurements of the casing radius were the CMM
1951: measurements of the inner radius. The outer surface of the casing
1952: was determined from the CMM values combined with the micrometer
1953: data.
1954:
1955:
1956:
1957:
1958: \subsection{Density Constraint}
1959: \label{sec46}
1960: Since the mercury represented roughly 90\% of the
1961: total tank mass, special attention was given to its contribution
1962: to the signal. The density of mercury samples from each tank was
1963: measured at the Physikalisch-Technische Bundesanstalt,
1964: Braunschweig with an accuracy of 3 ppm. One can use this density
1965: and mass measurements of the mercury (see Sec.~\ref{sec43}) to
1966: obtain better values for the effective tank dimensions and thus
1967: for the mass-integration constant. This results in a correlation
1968: among the effective parameters, $r,R,h,m$. The method employed to
1969: determine the best parameters representing the effective values
1970: (determined from measurements as described in the previous
1971: section) is based on minimizing a $\chi^2$ function of the form
1972: \begin{align}
1973: \chi^2=&\left(\frac{r-r_0}{\sigma_r}\right)^2
1974: +\left(\frac{R-R_0}{\sigma_R}\right)^2
1975: +\left(\frac{h-h_0}{\sigma_h}\right)^2 \notag \\
1976: &+\left(\frac{m-m_0}{\sigma_m}\right)^2
1977: +\left(\frac{\rho-\rho_0}{\sigma_\rho}\right)^2
1978: \end{align}
1979: subject to the density constraint
1980: \begin{equation}
1981: \rho=\frac{m}{\pi (R^2-r^2)h}.
1982: \end{equation}
1983: After substituting $\rho$ from Eq. (4) in Eq. (3), $\chi^2$
1984: becomes a function of the four parameters $r$, $R$, $h$, $m$, the
1985: five measured quantities $r_0$, $R_0$, $h_0$, $m_0$, $\rho_0$ and
1986: their uncertainties (see Table~\ref{tab:fitval} for the
1987: uncertainties of the effective values). The simplex method was
1988: used to minimize $\chi^2$ and thereby obtain best values for the
1989: fit parameters. Although $\rho$ is not explicitly one of the fit
1990: parameters, a best value for $\rho$ can be obtained by
1991: substituting best fit parameters in Eq. (4).
1992:
1993: The difference between best fit parameters and the measured values
1994: are shown in Table~\ref{tab:fitval} along with the resulting
1995: minimum $\chi^2$. It is seen that the difference between
1996: parameters and measured values is less than the error in all cases
1997: and that $\chi^2$ is consistent with the expected $\chi^2$ for a
1998: least-squares fit with one DOF.
1999:
2000: %------table 4 difference between measured and improved values-----
2001: \begin{table}[tbp]
2002: \caption{The correlated measured values, their uncertainties and
2003: the difference between the best fit parameters and the measured
2004: values for upper and lower tanks labelled 1 and 2. Only
2005: approximate values are shown for the measured quantities.}
2006: \label{tab:fitval}
2007: \begin{ruledtabular}
2008: \begin{tabular}{l r l r l r l r l}
2009: \multicolumn{1}{l}{}&
2010: \multicolumn{2}{c}{Measured}& \multicolumn{2}{c}{Uncertainty}&
2011: \multicolumn{2}{c}{Difference 1}&
2012: \multicolumn{2}{c}{Difference 2}\\
2013: \hline $r_0$ &60 &mm & 1.1 & $\mu \mbox{m}$ &0.007 & $\mu \mbox{m}$&0.013&$\mu \mbox{m}$
2014: \\$R_0$ &498 &mm & 9.0 & $\mu \mbox{m}$ &-3.5 &$\mu \mbox{m}$&-7.1&$\mu \mbox{m}$
2015: \\$h_0$ &650 &mm & 24.0 & $\mu \mbox{m}$ & -9.5 &$\mu \mbox{m}$&-19.0&$\mu \mbox{m}$
2016: \\$m_0$ &6760 &kg & 14.8 & g & 0.3 &g&0.7&g
2017: \\$\rho_0$ &13.54&g/cm$^{3}$ & 40.6 & $\mu \mbox{g/cm}^{3}$ &-1.3 &$\mu \mbox{g/cm}^{3}$&-2.5&$\mu \mbox{g/cm}^{3}$
2018: \\ \hline $\chi^2$&&&&&0.27&&1.18&
2019: \end{tabular}
2020: \end{ruledtabular}
2021: \end{table}
2022:
2023:
2024: In order to obtain the uncertainty of the mass-integration signal
2025: one needs the parameter-error matrix involving $r$, $R$, $h$, $m$
2026: and $\rho$ multiplied by partial derivatives of the signal with
2027: respect to the these quantities. The partial derivative with
2028: respect to $\rho$ is zero since it does not occur explicitly in
2029: the expression for the signal. The error of the signal $S$ is
2030: given by
2031: \begin{displaymath}
2032: \sigma_S=\left( \sum_{i,j}\frac{\partial S}{\partial x_i}
2033: \frac{\partial S}{\partial x_j} \textrm{err}(x_i,x_j) \right)^{1/2}
2034: \end{displaymath}
2035: where the $x_i$ and $x_j$ are any pair of the measured quantities
2036: and err$(x_i,x_j)$ is the 5 by 5 parameter-error matrix. Assuming
2037: that, for small variations about the measured values, the fit
2038: parameters represented by the 5-dimension vector $\overline y$ can
2039: be expressed as a linear function of the measured values
2040: $\overline x$ of the form $\overline y=T \overline x +\overline
2041: a$. The parameter-error matrix can be written as
2042: \begin{displaymath}
2043: \textrm{err}(y_i,y_j)=T V T^t
2044: \end{displaymath}
2045: where T is the Jacobi derivative of $y$ with respect to $x$, $T^t$
2046: is its transpose, $V$ is the 5 by 5 matrix covariance matrix (i.e.
2047: $V_{i,j}=err(y_i,y_j)$) with all zero elements except along its
2048: diagonal and $\overline a$ is a constant vector. The elements of
2049: the matrix $T$ and the vector $\overline a$ were determined
2050: numerically by solving a system of linear equations in which the
2051: fit parameters were determined for measured values incremented by
2052: small amounts ($\sigma_x$).
2053:
2054: The partial derivatives were also determined numerically by
2055: calculating the signal for measured values with small increments
2056: ($\sigma_x$). The signal was calculated using actual dimensions of
2057: the deformed tanks corrected by factors relating the $r,R,h$
2058: parameters to the effective dimensions $r_0,R_0,h_0$. The
2059: resulting covariance matrix representing the square of the
2060: uncertainty in the calculated mass-integration signal is shown in
2061: Table~\ref{tab:covar}. It is seen that the elements involving $R$
2062: and $h$ are much larger than those for $r$ and $m$. For the upper
2063: tank (tank 1), the relative uncertainty of the calculated
2064: mass-integration constant due to the correlated dimensions is
2065: 2.14~ppm. For tank 2, it is 2.41~ppm. The large cancellation
2066: occurring in the sum of the elements results in the uncertainty
2067: for these constrained parameters being approximately a factor of
2068: seven smaller than the uncertainty that would be obtained without
2069: the density information.
2070:
2071:
2072:
2073: %--------table 5 signal covariance matrix--------------
2074: \begin{table}
2075: \begin{ruledtabular}
2076: \caption{The covariance matrix involving $r,R,h,m$ for the
2077: uncertainty of the calculated signal. Values are given in units of
2078: ng$^2$. Only the upper part of the symmetric matrix is shown. The
2079: sum of all element in the full matrix is 2.84~$\mbox{ng}^2$. The
2080: sum of a similar diagonal matrix for uncorrelated $r,R,h,m$ values
2081: (not shown) is 175 $\mbox{ng}^2$.} \label{tab:covar}
2082:
2083: \begin{tabular}{l uuuu}
2084: & \mone{r} & \mone{R} & \mone{h} & \mone{m} \\ \hline
2085: r & 0.52 & 0.04 & 0.05 & 0.0003\\
2086: R & \mone{} & 35.77 & -42.52 & -0.20\\
2087: h & \mone{} & \mone{} & 51.62 & -0.25\\
2088: m & \mone{} & \mone{} & \mone{} & 0.66 \\
2089: \end{tabular}
2090: \end{ruledtabular}
2091: \end{table}
2092:
2093:
2094:
2095:
2096:
2097:
2098: \subsection{The Effect of Air Density}
2099: \label{sec47} Since air is not present in the region of the FM's,
2100: the motions of the FM's results in a force on the TM's due to the
2101: mass of the air elsewhere. It is as if there were a negative
2102: contribution to the mass of the FM's due to the lack of air in
2103: this region. This effect depends upon the density of the air in
2104: the region surrounding the FM's.
2105:
2106: The air pressure, the relative humidity and the air temperature
2107: were recorded every 12~min during the gravitational measurement
2108: thereby providing the information necessary to determine the air
2109: density. The effect of air density on the calculated
2110: mass-integration constant was approximately 100 ppm. The variation
2111: of the mass-integration constant for this effect was only about 1
2112: ppm. Thus, it was sufficient to employ only the average value of
2113: the air density during the entire gravitational measurement. The
2114: average density employed was 1.156~$\mbox{kg}/\mbox{m}^3$.
2115:
2116: \subsection{The Effect of Mercury Expansion}
2117: \label{sec48}
2118: Due to the small temperature variations of the FM's,
2119: the volume of the mercury relative to the volume of the tanks
2120: changed slightly during the gravitational measurement. This
2121: resulted in a variation of the mercury height in the compensation
2122: vessels. The height of the mercury in each compensation vessel was
2123: recorded every 12~min during the experiment. The calculated
2124: mass-integration constant varied by only 0.3~ppm due to this
2125: effect. Therefore, only an average value of the mercury height in
2126: each compensation vessel was employed in determining the
2127: mass-integration constant for the entire measurement.
2128:
2129:
2130: \subsection{Uncertainties Affecting the Mass-Integration Constant}
2131: \label{sec49}
2132: The relative uncertainties of the mass-integration
2133: constant due to the various measured and estimated quantities
2134: relating to either the upper or lower TM or to either the upper or
2135: lower FM are given in Table~\ref{tab:umassint}. The signs of the
2136: estimated quantities have been chosen to give the largest
2137: uncertainty of the mass-integration constant. With the exception
2138: of the constrained quantities $r$, $R$, $h$, and $m$, all measured
2139: quantities of this table are independent (i.e. uncorrelated). All
2140: estimated quantities are also independent. The total uncertainty
2141: of the mass-integration constant due to the measured quantities
2142: listed in Table~\ref{tab:umassint} results in a relative
2143: statistical uncertainty of 4.89~ppm. The total uncertainty of the
2144: mass-integration constant due to estimated quantities results in a
2145: relative systematic uncertainty of 3.25~ppm.
2146:
2147: %----------table 6 uncertainty mass integration constant---------
2148: \begin{table}
2149: \caption{Mass-integration constant relative uncertainties (ppm)
2150: associated with the measured quantities. 'Upper' and 'Lower' refer
2151: to the upper and lower FM or TM quantities. The values in
2152: parentheses are the uncertainty of the measured quantities. Where
2153: two measured values are listed, the first applies to the upper
2154: object and the second to the lower object. Quantities marked with
2155: a $^*$ are obtained from estimated limiting values. All
2156: uncertainties are independent except for the constrained
2157: quantities r,R,h,m. However, these constrained quantities are
2158: independent for the upper and lower tanks.} \label{tab:umassint}
2159: \begin{ruledtabular}
2160: \begin{tabular}{l ss}
2161: \mone{Measured Quantity} & \mone{Upper} &
2162: \mone{Lower} \\
2163: \hline & & \\FM Quantities & &
2164: \\ \hspace{5 mm} $r,R,h,m$ constrained & 1.20 & 1.09
2165: \\ \hspace{5 mm} position=2460 or 1042 mm (35 \mum) & 2.05 &
2166: 2.99
2167: \\ \hspace{5 mm} inner radius=50 mm (1.1 \mum) & 0.91 & 0.91
2168: \\ \hspace{5 mm} travel=709 mm (10 \mum) & 0.95 & 1.22
2169: \\ \hspace{5 mm} upper plate mass=153 kg (0.9~g) & 0.01 & 0.01
2170: \\ \hspace{5 mm} lower plate mass=154 kg (0.45~g) &0.02 & 0.02
2171: \\ \hspace{5 mm} central piece mass=46 kg (0.18~g) & 0.03 & 0.03
2172: \\ \hspace{5 mm} outer tube mass=412 kg (0.83~g) & 0.01 & 0.01
2173: \\ \hspace{5 mm} central piece no density gradient & 0.03 & 0.03
2174: \\ \hspace{5 mm} central piece z density gradient$^*$ &<0.03&<0.03
2175: \\ \hspace{5 mm} central piece r density gradient$^*$ &<1.1&<1.1
2176:
2177: \\ & &
2178: \\TM Quantities & &
2179: \\ \hspace{5 mm} radius=23 mm (5 \mum) & 0.57 & 0.57
2180: \\ \hspace{5 mm} height=77 mm (5 \mum) & 0.49 & 0.87
2181: \\ \hspace{5 mm} position=2495, 1077 mm (35 \mum) & 0.45 & 0.32
2182: \\ \hspace{5 mm} mass=1.1 kg (300 \mug) & 0.14 & 0.14
2183: \\ \hspace{5 mm} off center=0.44 or 0.51 mm (0.1 mm)&1.03 & 1.03
2184: \\ \hspace{5 mm} angle relative to vertical $^*$ &<1.85 &<1.85
2185: \\ \hspace{5 mm} relative z density gradient$^*$ &<0.9&<0.7
2186: \\ \hspace{5 mm} relative r density gradient$^*$ &<0.02&<0.02
2187: \\ \end{tabular}
2188: \end{ruledtabular}
2189: \end{table}
2190:
2191:
2192:
2193: In addition to the uncertainties related to either TM alone, there
2194: is the common vertical displacement (shown as $\varepsilon$ in
2195: Fig.~\ref{fig:geom}) of both TM's due to evacuating the system.
2196: The uncertainty of this displacement results in a relative
2197: uncertainty of the mass-integration constant equal to 0.78~ppm
2198: which is added to the other uncertainties as an independent
2199: relative uncertainty. Including the uncertainty of $\varepsilon$
2200: results in a relative statistical uncertainty of the
2201: mass-integration constant equal to 4.95~ppm.
2202:
2203:
2204: %****************Discussion of Measurements**************
2205: \section{Discussion of Measurements}
2206: \label{sec5} The measured gravitational signal discussed in
2207: Secs.~\ref{sec37} to \ref{sec311} is 784,883.3(12.2)(5.8)~ng. The
2208: calculated mass-integration constant determined in
2209: Secs.~\ref{sec41} to \ref{sec45} is 784,687.8(3.9)(2.6)~ng. Using
2210: these values, we obtain the value for the gravitational constant
2211: \\ \\
2212: $G$=6.674252(109)(54)$\times 10^{-11}\;\Gunit$.
2213: \\ \\
2214: A summary of the relative uncertainties contributing to this
2215: result is given in Table~\ref{tab:relerrs}.
2216:
2217: %-----------table 7, summary of errors-----------
2218: \begin{table}
2219: \caption{Relative statistical and systematic uncertainties of $G$
2220: as determined in this experiment.} \label{tab:relerrs}
2221: \begin{ruledtabular}
2222: \begin{tabular}{l l r}
2223: {Description}& {Statistical(ppm)} &{Systematic(ppm)}
2224: \\ \hline
2225: Measured Signal & &\\
2226: \hspace{5 mm} Weighings & 11.6 & \\
2227: \hspace{5 mm} TM-sorption & 7.4 & 7.4\\
2228: \hspace{5 mm} Linearity &6.1& \\
2229: \hspace{5 mm} Calibration & 4.0 & 0.5\\
2230: Mass Integration & 5.0&3.3\\
2231: \hline
2232: Total & 16.3 & 8.1 \\
2233: \end{tabular}
2234: \end{ruledtabular}
2235: \end{table}
2236: The relative statistical and systematic uncertainties of this
2237: result are 16.3~ppm and 8.1~ppm, respectively. The two largest
2238: contributions to the total relative uncertainty are the
2239: statistical uncertainty of the weighings (11.6~ppm) and the
2240: combined statistical and systematic uncertainty due to the
2241: TM-sorption effect (10.3~ppm). All uncertainties have been given
2242: as one sigma uncertainties. Statistical and systematic
2243: uncertainties have been combined to give a total uncertainty by
2244: taking the square root of the sum of their squares.
2245:
2246:
2247: \subsection{Comparison with Our Previous Analysis}
2248: \label{51}
2249: Our previously published value \cite{Sa02} for $G$ was
2250: $6.674070(220) \times 10^{-11}\; \Gunit$ which was based on the
2251: measurements of both the copper and tantalum TM's. The value for
2252: the copper TM's alone was $6.674040(210)\times 10^{-11}\;\Gunit$.
2253: The value obtained for $G$ in the present analysis for only the
2254: copper TM's ($6.674252((124)\times 10^{-11}\; \Gunit$) is in
2255: reasonable agreement with the previous value. The difference
2256: between the present and previous result is due primarily to the
2257: correction for the ZP curvature which was not taken into account
2258: in the previous analysis. A minor difference is also due to a
2259: slightly different selection of the analyzed data.
2260:
2261: The uncertainty given for the present result is appreciably
2262: smaller than that of our previous result. This is due to a better
2263: method used in computing the nonlinearity correction
2264: (Sec.~\ref{sec310}) and a calculation of the mass-integration
2265: constant (Sec.~\ref{sec45}) using the mercury density as a
2266: constraint. The uncertainty of the linearity correction was
2267: reduced from 18~ppm to 6.1~ppm and the uncertainty of the
2268: mass-integration constant was reduced from 20.6~ppm to 6.7~ppm.
2269: The statistical uncertainty of the weighings in the present
2270: analysis is somewhat larger than the previous value (9.1~ng vs
2271: 5.4~ng). This is due to the correlation of the ZP-corrected data
2272: of the present analysis. The previous ABA analysis involved only
2273: uncorrelated data.
2274:
2275: \subsection{Comparison with Other Measurements}
2276: \label{sec52}
2277: Recent measurements \cite{Gu00,Qu01,Ar03} of the
2278: gravitational constant with relative errors less than 50~ppm are
2279: listed in Table~\ref{tab:Gcomp}\ and shown in
2280: Fig.~\ref{fig:Gcomp}. We list only the latest publication of each
2281: group. It is seen that the present result is in good agreement
2282: with those of Gundlach and Merkowitz \cite{Gu00} and Fitzgerald
2283: \cite{Ar03}. It is in disagreement (3.6 times the sum of the
2284: uncertainties) with the result of Quinn et al. \cite{Qu01}.
2285:
2286: %-----------fig 20 plot of other measurements-------------
2287: \begin{figure}[width=8cm]
2288: \includegraphics[width=8.5cm]{G_measurements.ps}
2289: \caption{Plot of recent measurements with relative errors less
2290: than 50~ppm} \label{fig:Gcomp}
2291: \end{figure}
2292:
2293: All of the measurements listed in Table~\ref{tab:Gcomp} with the
2294: exception of our own were performed using torsion balances. It is
2295: therefore instructive to compare the problems encountered in the
2296: different types of measurements.
2297:
2298: In the measurements being discussed, the statistical accuracy in
2299: determining the gravitational signal was obtained in measurements
2300: lasting one to six weeks. However, as in the case of all precision
2301: measurements, the time required for obtaining a good statistical
2302: accuracy of the measurement is less than the time required to
2303: obtain calibration accuracy of the equipment and the time
2304: necessary to investigate and eliminate systematic errors. All of
2305: the measurements listed in Table~\ref{tab:Gcomp} have been long, on-going
2306: investigations which have lasted for periods up to ten years.
2307:
2308: Although the beam-balance measurement was made with more massive
2309: FM's (15~t) than were employed in the torsion-balance measurements
2310: ($<60$~kg), the larger gravitational signal had to be measured in
2311: the presence of the TM weight. In our experiment the gravitational
2312: signal was roughly 0.7~ppm of the total weight on the balance.
2313: This small ratio of signal to total weight on the balance resulted
2314: in larger effects of zero-point drift as well as larger
2315: statistical noise in the beam-balance data than in the
2316: torsion-balance data. In the torsion-balance measurements, the
2317: deflection of the balance arm is due entirely to the gravitational
2318: force to be determined with only small perturbations from distant
2319: moving objects.
2320:
2321: A similar problem has to do with the change of TM weight that is
2322: produced by an adsorbed water layer. In our measurement, this
2323: varied with the temperature of the vacuum tube produced by the FM
2324: motion. This resulted in one of the largest contribution to the
2325: uncertainty of the gravitational signal (see Sec.~\ref{sec310}).
2326: In the beam-balance measurement the weight change adds directly to
2327: the signal whereas in the torsion-balance it adds only to the mass
2328: of the torsion-arm and is therefore a negligible effect.
2329:
2330: The calibration of the beam balance, while simple in principle, is
2331: difficult in practice due to the lack of calibration masses having
2332: the required mass and accuracy. We have used an averaging method
2333: to correct for the nonlinearity of the balance (see
2334: Sec.~\ref{sec311}) involving a large number of auxiliary masses.
2335: This allowed the comparison of the gravitational signal with a
2336: heavier, accurately known calibration mass. The accurate
2337: calibration of the torsion balance also presents a problem.
2338: Various methods involving either electric forces or an angular
2339: acceleration of the measuring table to compensate the
2340: gravitational force have been used.
2341:
2342:
2343:
2344: In our measurement, the TM was positioned at a double extremum of
2345: the force field produced by the FM's. This greatly reduces the
2346: relative accuracy of the distance measurements required to
2347: determine the mass-integration constant. It also reduces the
2348: problem resulting from a density gradient in the TM. It is
2349: difficult to compare the problems involved in determining the
2350: mass-integration constant for the two types of experiments. It
2351: appears that the distance between the field masses attracting the
2352: small mass of the torsion balance must be measured with very high
2353: absolute accuracy (1~\mum\ is the accuracy given for this
2354: distance in the torsion balance experiments).
2355:
2356: The use of liquid mercury as the principle component of the FM's
2357: reduces the problem associated with the density gradients of the
2358: FM's. There is still the density gradient of the vessel walls
2359: which has to be considered. The field masses employed in the
2360: torsion-balance experiments were either spheres or cylinders. The
2361: FM's were rotated between measurements in order to compensate
2362: gradient effects.
2363:
2364: The large mercury mass resulted in deformations of the vessel
2365: which had to be accurately determined. The FM's although, nearly
2366: cylindrical in form, required more than 1000 ring and point-mass
2367: elements in order to determine the mass-integration constant. A
2368: similar problem (but on a smaller scale) occurs in the torsion
2369: experiments in accounting for small imperfections of the spheres,
2370: cylinders or plates and in determining their relative positions.
2371:
2372: The determination of $G$ using a beam balance is beset with a
2373: number of problems which we have tried to describe in detail. We
2374: have been able to reduce the uncertainty in $G$ resulting from
2375: these problems to values comparable to the statistical
2376: reproducibility of the weighings determining the gravitational
2377: signal. The total uncertainty for $G$ which we obtain with a beam
2378: balance is comparable to the uncertainty quoted in recent
2379: torsion-balance determinations of $G$. We believe that the
2380: beam-balance measurement involving a number of quite different
2381: problems than encountered in torsion balance measurements can
2382: therefore provide a useful contribution to the accuracy of the
2383: gravitational constant.
2384:
2385:
2386:
2387:
2388: %--------------table 8 other measurements----------------
2389: \begin{table}[tbp]
2390: \caption{Recent measurements of the gravitational constant.
2391: Statistical and systematic uncertainties have been added as if the
2392: were independent quantities.} \label{tab:Gcomp}
2393: \begin{ruledtabular}
2394: \begin{tabular}{l l}
2395: \multicolumn{1}{l}{Reference}&
2396: \multicolumn{1}{c}{$G[10^{-11}\;\Gunit]$}\\
2397: \hline
2398: Gundlach and Merkowitz \cite{Gu00} &6.674215( 92)\\
2399: Quinn et al.\cite{Qu01} &6.675590(270)\\
2400: Armstrong and Fitzgerald \cite{Ar03}&6.673870(270)\\
2401: Present analysis &6.674252(124) \\
2402: \end{tabular}
2403: \end{ruledtabular}
2404: \end{table}
2405:
2406:
2407: \section{Acknowledgements}
2408: \label{sec6}
2409: The present research has been generously supported by
2410: the Swiss National Fund, the Kanton of Z\"{u}rich, and
2411: assistantship grants provided by Dr. Tomalla Foundation for which
2412: we are very grateful. The experiment could not have been carried
2413: out without the close support of the Mettler-Toledo company which
2414: donated the balance for this measurement and made their laboratory
2415: available for our use. We are also extremely grateful to the Paul
2416: Scherrer Institut for providing a suitable measuring room and the
2417: help of their staff in determining the geometry of our experiment.
2418: We also wish to thank the Swiss Metrological Institute and the
2419: Physikalisch-Technische Bundesanstalt, Braunschweig, Germany for
2420: making a number of certified precision measurements for us. Our
2421: heartfelt thanks are also extended to the staff of our machine
2422: shop for their advice and precision work in producing many of the
2423: pieces for this experiment. We wish to thank E. Klingel\'{e} for
2424: determining the value of local gravity at the measuring site. We
2425: are also indebted to the firms Almaden Co., Metrotec A.G. and
2426: Reishauer A.G. for the services they provided.
2427:
2428:
2429:
2430: %***************Appendix************************
2431: \section{Appendix}
2432: \label{sec7} The general idea in determining the $z$ component of
2433: force on a small volume element of a TM is first to calculate the
2434: potential along the $z$ axis due to the FM. The off-axis potential
2435: can then be obtained by making a Taylor series expansion for small
2436: $r$ and substituting this in the Laplace equation. This is the
2437: procedure which is often used in electrostatic calculations
2438: \cite{Gl56}. The force in the $z$ direction is just the negative
2439: derivative of this potential with respect to $z$. We present first
2440: a derivation of the force for the axially symmetric case and then
2441: describe the modification required for a nonaxially symmetric
2442: potential.
2443:
2444: The gravitational potential on the $z$ axis of a homogeneous,
2445: torus of rectangular cross section with density $\rho_{FM}$, inner
2446: radius $R_1$, outer radius $R_2$ and half height $Z$ is given by
2447: the equation
2448: \begin{equation}
2449: \Phi(r=0,z) = 2\pi\rho_{FM} G \int_{-Z/2}^{Z/2} \int_{R1}^{R2}
2450: \frac{r' dr' \;dz'}{\sqrt{{r'}^2 + (z'- z)^2}}
2451: \label{eq:Phi}
2452: \end{equation}
2453: where $r$ and $z$ are the radial and axial coordinates of a point
2454: within the TM expressed in cylindrical polar coordinates. For
2455: convenience, one chooses the zero of the potential at the center
2456: of mass of a TM. For simplicity, the cross section of the FM
2457: employed in Eq.~\ref{eq:Phi} has been chosen to be a rectangle.
2458: Besides the torus with rectangular cross section, analytic
2459: expressions for the two dimensional integrals with triangular and
2460: circular cross sections were also employed.
2461:
2462: The potential at points close to the axis can be calculated as a
2463: power series in $r$
2464: \begin{equation}
2465: \Phi(r,z) = a_0 + a_1 r + a_2 r^2 + a_3 r^3 + \ldots =
2466: \sum_{i=0}^\infty a_ir^i \;
2467: \label{eq:series}
2468: \end{equation}
2469: with unknown coefficients $a_1$,$a_2$,$\ldots$ which are functions
2470: of $z$ alone. For $r$=0, one has $a_0=\Phi(0,z)$. The
2471: gravitational field satisfies the Laplace equation $\nabla^2 \Phi
2472: = 0$. Applying the Laplace operator in cylindrical coordinates to
2473: Eq.~\ref{eq:series} leads to
2474: \begin{displaymath}
2475: \nabla^2 \Phi=a_1r^{-1} + \sum_{i=0}^\infty r^i \left( \frac{d^2
2476: a_i}{d z^2} + \left(i+2 \right)^2 a_{i+2} \right) = 0 \; .
2477: \end{displaymath}
2478: This equation is valid for all $r$. Since $\nabla^2 \Phi=0$ for
2479: $r=0$, $a_1$ must be identically zero. Thus, the values for $a_i$
2480: can be calculated recursively from
2481: \begin{displaymath}
2482: a_{i+2} = - \frac{1}{\left( i+2 \right)^2} \frac{d^{2}
2483: a_i}{dz^{2}}
2484: \end{displaymath}
2485: starting with either $a_0$ or $a_1$. Since $a_1=0$, all terms with
2486: odd $i$ are zero and $a_i$ for even $i$ can be obtained starting
2487: with $a_0$. By induction, it is easily shown that the coefficient
2488: $a_{2n}$ is
2489: \begin{displaymath}
2490: a_{2n} = \left( - \frac{1}{4} \right)^n \frac{1}{\left( n!
2491: \right)^2} V^{(2n)}(z)
2492: \end{displaymath}
2493: with
2494: \begin{displaymath}
2495: V^{(2n)}(z)=\frac{ d^{2n} \Phi(0,z) }{ d z^{2n} } .
2496: \end{displaymath}
2497: Here, $V^{(0)}$ is just $\Phi(0,z)$ which can be easily calculated
2498: using Eq.~\ref{eq:Phi}.
2499:
2500: Using this expression for the coefficients in the
2501: expansion~\ref{eq:series},
2502: the gravitational potential can be calculated from the sum
2503: \begin{equation}
2504: \Phi(r,z) = \sum_{i=0}^\infty \left( - \frac{1}{4} \right)^i
2505: \frac{1}{\left( i! \right)^2} V^{(2i)}(z)\; r^{2i}
2506: \label{eq:potential_sum} \;.
2507: \end{equation}
2508:
2509:
2510: The gravitational field $g_z$ in z-direction is given by
2511: $-\partial \Phi / \partial z$ or
2512: \begin{displaymath}
2513: g_z(r,z) = -\sum_{i=0}^\infty \left( - \frac{1}{4} \right)^i
2514: \frac{1}{\left( i! \right)^2} V^{(2i+1)}(z) \; r^{2i} \; .
2515: %\label{eq:field_sum}
2516: \end{displaymath}
2517: %--------------eq 8---------------------
2518: Integrating $g_z$ over the volume of the TM, one obtains the force
2519: on the TM in the $z$ direction
2520: \begin{align}
2521: &F_z =-2 \pi \rho_{TM} \times \notag \\
2522: &\sum_{i=0}^\infty \left( - \frac{1}{4}
2523: \right)^i \frac{1}{\left( i! \right)^2} \int_{-b}^{+b} \ d z'
2524: \int_0^r V^{(2i+1)}(z) \; r'^{2i} r' \ d r' \;
2525: \label{eq:F_sum}
2526: \end{align}
2527: where the origin has been taken to be the center of the TM. The
2528: integration over $r$ is trivial. The integration over $z$ is
2529: \begin{displaymath}
2530: \int_{-b}^{+b} V^{(2i+1)} d z' = V^{(2i)}(b) - V^{(2i)}(-b) \;
2531: .\nonumber
2532: \end{displaymath}
2533:
2534: Making a Taylor expansion for small $b$ on the right side of
2535: equation, one obtains
2536: \begin{eqnarray}
2537: \lefteqn{\int_{-b}^{+b} V^{(2i+1)} d z' =} \nonumber \\
2538: & V^{(2i)}(0) + b V^{(2i+1)}(0) + \frac{1}{2!} b^2 V^{(2i+2)}(0)
2539: + \ldots
2540: \nonumber \\
2541: & -V^{(2i)}(0) + b V^{(2i+1)}(0) - \frac{1}{2!} b^2
2542: V^{(2i+2)}(0) + \ldots \nonumber \; .
2543: \end{eqnarray}
2544: Adding similar terms results in
2545: %-------------eqs 9--------------------
2546: \begin{align*}
2547: \int_{-b}^{+b} V^{(2i+1)} d z' = & 2 b V^{(2i+1)}(0) +
2548: \frac{2}{3!} b^3 V^{(2i+3)}(0)\\ \notag
2549: & + \frac{2}{5!}b^5 V^{(2i+5)}(0) + \ldots \notag
2550: \end{align*}
2551: or
2552: \begin{equation}
2553: \int_{-b}^{+b} V^{(2i+1)} d z' =\sum_{j=0}^\infty
2554: \frac{2}{(2j+1)!} b^{2j+1} V^{(2i+2j+1)}(0) \; .
2555: \label{eq:integral_sum}
2556: \end{equation}
2557:
2558: The final equation for the gravitational force on a cylinder can
2559: then be calculated by combining Eq.~\ref{eq:F_sum} and
2560: \ref{eq:integral_sum} to obtain
2561: %--------------eqs 10------------------------------
2562: \begin{align}
2563: & F_z=-2 \pi b r^2 \rho_{TM} \notag \sum_{n=0}^\infty V^{(2n+1)}(0) \times \\
2564: &
2565: \sum_{i=0}^n \frac{1}{\left( -4 \right)^i}
2566: \frac{1}{i!\left(i+1\right)!} \frac{1}{\left( 2 n- 2 i +1
2567: \right)!} b^{ 2 n -2 i } r^{2i} \ .
2568: \label{eq:massint_main}
2569: \end{align}
2570:
2571: Only odd derivatives of the potential are required for this case
2572: involving complete axial symmetry. The convergence of the series
2573: can be improved by dividing the TM into two or more shorter
2574: cylinders. The algebraic expressions for the $V^{2n}(0)$, as
2575: determined using "automatic differentiation" \cite{Ra81}, are very
2576: long and will not be given here.
2577:
2578: The above derivation has assumed that the TM and FM have a common
2579: axis of cylindrical symmetry. We will now show how essentially the
2580: same equations can be employed for an arbitrary FM potential. This
2581: allows one to calculate the potential of a FM with cylindrical
2582: symmetry but with its axis displaced and/or tilted relative to
2583: that of the TM.
2584:
2585: Again, in order to facilitate integration over the volume of the
2586: cylindrical TM, one employs cylindrical polar coordinates with the
2587: $z$ axis along the symmetry axis of the TM. The potential is now
2588: a potential of the form $\psi(x,y,z)=\psi\left (r\cos(\varphi),
2589: r\sin(\varphi),z\right)$ which satisfies $\nabla^2\psi=0$. The
2590: center of mass of the TM is chosen as the zero of potential.
2591:
2592: One defines a function $\Psi(r,z)$ such that
2593: \begin{equation}
2594: \Psi(r,z)=(2 \pi)^{-1}\int_{\varphi=0}^{2\pi} \psi(r,\varphi,z)
2595: d\varphi.
2596: \label{eq:Psi}
2597: \end{equation}
2598: One can then show that this function satisfies the same assumption
2599: that were made for the function $\Phi(r,z)$, namely that $\Psi$
2600: has axial symmetry, that it is zero on the $z$ axis and that its
2601: Laplacian is zero. It can therefore be used in the Eqs. 7 through
2602: 10 instead of $\Phi$ to obtain the gravitational force integrated
2603: over the TM.
2604:
2605: The axial symmetry of $\Psi$ is obvious since $\varphi$ has been
2606: removed by the integration over $\varphi$. The zero potential was
2607: chosen to be at the TM center of mass. The Laplacian of $\Psi$ can
2608: be shown to be zero by allowing the Laplacian to operate on $\Psi$
2609: as defined in Eq.~\ref{eq:Psi}. Reversing the integration and
2610: differentiation operations one obtains
2611: \begin{displaymath}
2612: \nabla^2 \Psi(r,z) = (2 \pi)^{-1} \int_{\varphi=0}^{2
2613: \pi} (\nabla^2 \psi-r^{-2}\frac{\partial^2 \psi}{\partial \varphi^2})
2614: d\varphi.
2615: \end{displaymath}
2616: The Laplacian of $\psi(r,\varphi,z)$ is zero for an inverse $r$
2617: potential. The integral of the second term is $r^{-2}\partial \psi
2618: / \partial \varphi$ evaluated at $\varphi=0$ and $2\pi$ which is
2619: also zero. Thus, one obtains the desired result that $\nabla^2
2620: \Psi(r,z)=0$.
2621:
2622: In order to use this property of a potential which does not have
2623: axial symmetry about the TM axis (the $z$ axis), one must
2624: determine the potential and its derivative with respect to $z$
2625: along the $z$ axis. This is not difficult for the case of a FM
2626: which has axial symmetry about an axis not coincident with the TM
2627: axis. One merely uses Eq. 7 to determine the potential at radial
2628: distances from the FM axis corresponding to points on the TM axis.
2629: The force in the $z$ direction (TM axis) is then determined as
2630: before using Eqs. 8 and 10. This procedure is particularly useful
2631: for the case of TM and FM axes which are parallel but which are
2632: displaced relative to one another.
2633:
2634: In principle, one can determine the derivatives with respect to
2635: $r$ which are required for the force on a TM tilted relative to
2636: the vertical; however, in this case it is simpler to approximate
2637: the TM by a number of thin slabs displaced from the vertical axis.
2638: This completes the discussion of nonaxial-symmetric potentials.
2639:
2640: \begin{thebibliography}{99}
2641: \bibitem{Mo05} P.~J.~Mohr and B.~N.~Taylor, Rev. Mod. Phys. 77, 1(2005).
2642: \bibitem{Fi86}
2643: E.~Fischbach, D.~Sudarsky, A.~Szafer, C.~Talmadge and S.~H.~Aronson,
2644: Phys.\ Rev.\ Lett.\ 56, 3(1986)
2645: [Erratum-ibid.\ 56, 1427(1986)].
2646: \bibitem{Co94} A.~Cornaz, B.~Hubler and W.~K\"{u}ndig, Phys. Rev. Lett. 72, 1152(1994).
2647: \bibitem{Hu95} B.~Hubler, A.~Cornaz and W.~K\"{u}ndig, Phys. Rev. D51, 4005(1995).
2648: \bibitem{Mo88} G.~I.~Moore, F.~D.~Stacey, G.~J.~Tuck, B.~D.~Goodwin,
2649: N.~P.~Linthorne, M.~A.~Barton, D.~M.~Reid and G.~D.~Agnew, Phys.
2650: Rev. D38, 1023(1988).
2651: \bibitem{Sp87} C.~C.~Speake and G.~T.~Gillies, Proc. R. Soc.
2652: London A414, 415(1987).
2653: \bibitem{Sa98}J.~Schurr, F.~Nolting and W.~K\"{u}ndig, Phys. Lett. A248, 295(1998).
2654: \bibitem{Sb98}J.~Schurr, F.~Nolting and W.~K\"{u}ndig, Phys. Rev. Lett. 80, 1142(1998).
2655: \bibitem{No98}F.~Nolting, PhD Thesis, University of Zurich(1998).
2656: \bibitem{No99}F.~Nolting, J.~Schurr, St.~Schlamminger and W.~K\"{u}ndig,
2657: IEEE Trans. Instrum. Meas. 48, 245(1999).
2658: \bibitem{Sa99}F.~Nolting, J.~Schurr, St.~Schlamminger and W.~K\"{u}ndig,
2659: Meas. Sci. Technol. 10, 487(1999).
2660: \bibitem{Sa02}St.~Schlamminger, E.~Holzschuh and W.~K\"{u}ndig,
2661: Phys. Rev. Lett. 89,161102-1(2002).
2662: \bibitem{Sb02}St.~Schlamminger, PhD Thesis, University of Zurich(2002).
2663: \bibitem{Sl88}A.~H.~Slocum, Precision Eng. 10(2), 85(1988).
2664: \bibitem{Fu81}J.~E.~Furse, J. Phys., E:Sci. Instrum. 14, 264(1982).
2665: \bibitem{Sch94a}R.~Schwartz, Metrologia, 31, 117(1994).
2666: \bibitem{Sch94b}R.~Schwartz, Metrologia, 31, 129(1994).
2667: \bibitem{Re00}A.~Reichmuth, Non-Linearity in Mass Comparison.
2668: Proc. Intl. Conf. Metrology, Jerusalem (II), pages 58-66,
2669: May(2000).
2670: \bibitem{Ba89} R.~J.~Barlow, Statistics, John Wiley and Sons, Chichester(1989).
2671: \bibitem{Gl56}W.~Glaser, Encyclopedia of Physics, Volume XXXIII,
2672: Springer-Verlag, Berlin-G\"{o}ttingen-Heidelberg(1956).
2673: \bibitem{Ti87} S.~P.~Timoshenko and S.~Woinowsky-Krieger, Theory of Plates and
2674: Shells, McGraw-Hill, New York(1987).
2675: \bibitem{Ra81}L.~B.~Rall, Lecture Notes in Computer Science 120:
2676: Automatic Differentiation: Techniques and Applications, edited by
2677: G.~Goos and J.~Hartmanis, Springer-Verlag, New York(1981)
2678: Publishers, Amsterdam(1989).
2679: \bibitem{Gu00}
2680: J.~H.~Gundlach and S.~M.~Merkowitz, Phys.\ Rev.\ Lett.\ 85, 2869
2681: (2000).
2682: \bibitem{Qu01} T.~J.~Quinn, C.~C.~Speake, S.~J.~Richman,
2683: R.~S.~Davis and A.~Picard, Phys. Rev. Lett. 87, 111101-4(2001).
2684: \bibitem{Ar03}T.~R.~Armstrong and M.~P.~Fitzgerald, Phys. Rev. Lett. 91, 201101-4(2003).
2685: \end{thebibliography}
2686:
2687: \end{document}
2688: