• Aucun résultat trouvé

Measurement of the Casimir force in a gas and in a liquid

N/A
N/A
Protected

Academic year: 2021

Partager "Measurement of the Casimir force in a gas and in a liquid"

Copied!
9
0
0

Texte intégral

(1)

HAL Id: hal-01830823

https://hal.archives-ouvertes.fr/hal-01830823v2

Preprint submitted on 7 Nov 2018

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

Measurement of the Casimir force in a gas and in a

liquid

Anne Le Cunuder, Artyom Petrosyan, Georges Palasantzas, Vitaly Svetovoy,

Sergio Ciliberto

To cite this version:

Anne Le Cunuder, Artyom Petrosyan, Georges Palasantzas, Vitaly Svetovoy, Sergio Ciliberto.

Mea-surement of the Casimir force in a gas and in a liquid. 2018. �hal-01830823v2�

(2)

Anne Le Cunuder1,∗, Artyom Petrosyan1, Georges Palasantzas2, Vitaly Svetovoy2,3 and Sergio Ciliberto1

1Laboratoire de Physique, CNRS UMR5672,

Université de Lyon, École Normale Supérieure, 46 Allée d'Italie, 69364 Lyon, France

2 Faculty of Sciences and Engineering, University of Groningen,

Nijenborg 4, 9747 AG Groningen, The Netherlands and

3 A. N. Frumkin Institute of Physical Chemistry and Electrochemistry,

Russian Academy of Sciences, Leninsky prospect 31 bld. 4, 119071 Moscow, Russia We present detailed measurements of the Casimir-Lifshitz force between two gold surfaces, per-formed for the rst time in both gas (nitrogen) and liquid (ethanol) environments with the same apparatus and on the same spot of the sample. Furthermore, we study the role of double-layer forces in the liquid, and we show that these electrostatic eects are important. The latter contributions are subtracted to recover the genuine Casimir force, and the experimental results are compared with calculations using Lifshitz theory. Our measurements demonstrate that carefully accounting for the actual optical properties of the surfaces is necessary for an accurate comparison with the Lifshitz theory predictions at distances smaller than 200nm.

Introduction. As devices enter the submicron range,

Casimir forces [112] between neutral bodies at close

proximity become increasingly important. As Casimir

rst understood in 1948 [2], these forces are due to the

connement of quantum uctuations of the electromag-netic (EM) eld. Indeed Casimir proved that when two parallel, perfectly reecting plates, are introduced in vac-uum, they impose, on the EM eld, boundary conditions which select only the uctuations compatible with them. As a result, an attractive force between the plates is pro-duced, which depends only on fundamental constants, on the distance d between the surfaces and on their area A:

Fc(d) = −

π2

A~c

240 d4 (1)

with ~ the Planck constant and c the speed of light.

Fol-lowing Casimir's calculation [2], Lifshitz and co-workers

in the 50's [3] considered the more general case of real

dielectric plates by exploiting the uctuation-dissipation theorem, which relates the dissipative properties of the plates and EM uctuations at equilibrium. Furthermore, for real surfaces, roughness and material optical

proper-ties can strongly alter the Casimir force [13,14].

Lifshitz formalism describes the Casimir force in a gen-eral case, where the medium between the plates need not be vaccum. According to this formalism, the force can be tuned from attractive to repulsive with a suitable choice of the interacting materials. These predictions boosted Casimir experiments to test the possibility of repulsive

forces [16]. In liquids, the determination of the Casimir

force is more complex than in a gas because of the pres-ence of additionnal eects, as e.g. the Debye screening. The Casimir-Lifshitz force has been measured between

two gold surfaces immersed in ethanol [17]; in this

exper-iment, electrostatic forces are found to be negligible as sodium iodide (NaI) was added to ethanol, decreasing the Debye screening length. However, the role of electrostatic forces and their screening by the Debye layer are

impor-É DES SURFACES

eux surfaces s’exercent à une hauteur moyenne comprise ent des aspérités. Il est donc indispensable de déterminer cette

connaître la distance de séparation effective pendant la m s surfaces rugueuses nécessite une approche statistique. L a ontact d0 est déterminée à partir de la distribution en ha

Thi s con di t i on can be con si der ed as an equat i on for t he h ei gh t of t h e h ei ght asper i t y due t o a shar p expon en t i al beh avi or of t he di st r i but i on

d0

smooth

a h1(x,y) h2(x,y)

orsque les deux surfaces rugueuses entrent en contact mé ts des aspérités de chacune des surfaces qui se touchent. L urs séparées par une distance de contact d > d0> 3w

ation la plus critique, le pic le plus élevé sur la surface de la encontre le pic le plus élevé (d0,ech) de la surface de l’éch ale de séparation est donc donnée par : d0,max= d0,ech+d0,sp.

e de se produire, sauf si les surfaces ont des rugosités comp la sphère est bien plus rugueuse que l’échantillon On con

d

t d

0

est déterminée à partir de la distribution en hauteur de la

Thiscondition can beconsidered asan equation for theheight of the ght asperity due to a sharp exponential behavior of the distribution

d

0

h

a

h

1

(dxddd,y)

h

2

(x,y)

dpiezo - h1 h2 d0=h1+h2 piezoelectric element d

FIG. 1. Experimental setup. A gold-coated polystyrene bead is glued on the cantilever tip which measures the sphere-plane interaction force at distance d. The deection of the cantilever is detected by an interferometric technique: two lasers beams, orthogonally polarized, are focused on the can-tilever, the reference one is reected by the static base and the second one by the cantilever free end. When the cantilever is bended the optical path dierence δ between the two beams is measured through an interferometer [15].

tant and one has to consider carefully their contributions

during force measurements in liquids [18,19]. In order to

clarify the interplay of the Casimir force and additionnal eects in liquids, we have performed measurements of the Casimir force in a nitrogen atmosphere in the rst place, and then, using the same system and sample, in ethanol. The contact area is the same in both measurements. We observe that electrostatic forces, screened over the Debye length, are of the same magnitude as the Casimir force, in the 50-200nm distance range. After subtracting the electrostatic force, we obtain a Casimir force in

quanti-tative agreement with Lifshitz theory [3]. Furthermore,

(3)

2 the importance of accurately characterizing the optical

properties of the samples before any meaningful compar-ison with theory.

Experimental setup. We use an atomic force micro-scope (AFM) to measure the Casimir force between metallic surfaces. In order to measure the force with a good accuracy, the cantilever displacement is measured

with a home-made quadrature phase interferometer [15],

whose operating principle is sketched on Fig.1.

The experiment is performed in a sphere-plane geom-etry to avoid the need to maintain two at plates per-fectly parallel. Thus, a polystyrene sphere of radius

R = (75.00 ± 0.25)µm (Sigma-Aldrich) is mounted on

the tip of the cantilever with a conductive glue and then the whole probe is coated by a gold lm whose thickness is about 100nm. The plates have been gold

coated using cathodic sputtering by ACM, at the

LMA-CNRS. The diameter of the sphere has been determined

from Scanning Electron Microscopy. We use a cantilever (size 500µm × 30µm × 2.7µm, NanoAndMore) of stiness

κ = 0.57 ± 0.03N/m. The precise value of κ is

deter-mined using equipartition, ie. < δ2 >= kBT

κ , where kB

is the Boltzmann constant and T the temperature. The resonance frequency of the sphere-cantilever ensemble is

fo= 2271Hzin vacuum.

The sphere faces a glass at plate which is coated by

a gold lm of a thickness of about 100nm [20].

Accord-ing to [21], the layer is suciently thick to be considered

as a bulk-like lm. The plate is mounted on a piezo actuator (PZ38, Piezojena) which allows us to control the plane-sphere distance. During the experiment, the plate is moved continuously towards the sphere and the induced deexion of the cantilever is detected by the in-terferometer.

In air, because of water vapor, the capillary force far exceeds the Casimir force. Therefore, our measurement in a gas was performed after lling our cell with nitrogen. Calibrations. The total force between the surfaces is

the sum of the Casimir force Fcas(d)and additionnal

con-tributions:

Ftotal= Fcas(d) + Fel(d) + FH(d, v) . (2)

Electrostatic forces Fel(d)are due to a potential

dier-ence between surfaces, owing to dierdier-ences between the work functions of the materials used, and the possible

presence of trapped charges [22]. Hydrodynamic forces

FH(d, v) are due to the motion of the uid during the

approach of the plate towards the sphere, and depend

on their relative velocity v [23]. These hydrodynamic

ef-fects are negligible in a nitrogen atmosphere, where the

viscosity is γ = 1.76 10−6Pa s, but have to be

consid-ered in ethanol where the viscosity is 1000 times higher

(γ = 1.2 10−3Pa s).

There are two main requirements for a precise deter-mination of the Casimir force. Firstly, additionnal forces

must be measured with accuracy and subtracted from the total measured force. Secondly, because the force has a strong dependance on the distance between sur-faces, an independant measure of the distance is nec-essary, which becomes dicult when the separation ap-proaches nanometer scales. The diculty originates prin-cipally from surface roughness: when the two surfaces come into contact, the highest asperities of each surface touch each other and the surfaces are still separated by

a distance upon contact d0 [24].

The piezo actuator includes a position sensor which

gives us the displacement of the plate: dpiezo. We

de-ne the origin of dpiezo as the position of contact of the

highest peak of the sphere roughness with the surface of the plate, as the sphere is much rougher than the plate. The eective separation distance which appears in the

expression of the force can be written as (see Fig.1):

d = dpiezo+ d0− δ (3)

where d0 is the distance upon contact due to surface

roughness and δ is an additional correction which results from the static deexion of the cantilever in response to the total force Ftotal.

We determined the separation upon contact d0 from

hydrodynamic calibration, performed in ethanol. Imme-diatly after measuring the Casimir force in nitrogen, we injected carefully ethanol into the cell, and we performed calibrations and measurements of the Casimir force in ethanol. As the horizontal drift of our system is negli-gible, the contact area and the separation distance upon

contact d0 are the same in each measurement. This

as-sumption is further justied a posteriori: our experimen-tal curves all superimpose on top of each other and on top of the theoretical curve after shifting the distance by

the same value of d0. The hydrodynamic calibration is

presented in next section, while topographic analysis is

presented in [20]. The value of d0obtained from

hydrody-namic calibration is comparable with the value obtained from roughness analysis.

Hydrodynamic calibration in ethanol. The theoretical expression of the hydrodynamic force, for non-slip

bound-ary conditions, is given by [23]:

FH= −

6πηR2

d v (4)

where η is the uid viscosity, R is the radius of the sphere,

and v = ∂d

∂t is the relative velocity between the plate

and the sphere. Indeed, in our case, the slippage can be neglected as the mean free path is in the order of inter-molecular distances (e.g a few Angstroms), and negligible in comparison with the roughness of the surface (e.g a few

tens of nanometers) [27]

As is clear from Eq. (2), among the dierent forces

(4)

100 150 200 250 300 350

Distance [nm]

140 120 100 80 60 40 20 0 20

Casimir Force [pN]

(a)

100 150 200 250 300 350

Distance [nm]

1.00 0.75 0.50 0.25 0.00 0.25 0.50 0.75 1.00

(F

exp.

F

th.

)/|

F

th.

|

(b)

FIG. 2. Measurement of the Casimir force between two Au-surfaces in a nitrogen atmosphere. (a) Blue points correspond to the mean measured force. Blue circles correspond to a single measurement of the Casimir force. Orange solid curve corresponds to the Lifshitz theory in which the dielectric function  is evaluated from measured optical data of a real gold lm [25]. For the green dashed-dotted line,  is evaluated using the handbook optical data [26]. Red dashed line corresponds to the theory in the case of ideal conductors. (b) Relative dierence between the theoretical and experimental Casimir forces. Same colors as in (a). The error bars include the statistical and the systematic errors due to uncertainties on the separation upon contact d0= (31 ± 2)nm(using the hydrodynamic estimation), on the stiness κ = (0.57 ± 0.03)N/m and on the diameter

of the sphere D = (150.0 ± 0.5)µm.

is the only one which depends on velocity. Thus we performed two force measurements, moving continuously the plate towards the sphere: a rst one at velocity

v1 = 348nm s−1, and a second one at a velocity v2 =

5109nm s−1. By taking the dierence between these two

measurements, we cancelled all the velocity-independent

forces and from Eq. (4) we obtained FH measured at

v = v2− v1= 4742nm s−1 :

Ftotal(d, v20) − Ftotal(d, v01) = FH(d, v) . (5)

Here, v0

2 and v01 are the relative velocities between the

sphere and the sample, which are not exactly the piezo

velocities v2 and v1 because the cantilever is deected

when the plate is moved towards the sphere. v0

2 and v01

were determined precisely measuring the deection of the cantilever.

Measurements of the hydrodynamic force are presented

in [20]. Comparing the measured hydrodynamic force

with the theoretical expression of Eq. (4), we determined

the separation distance upon contact d0= (31 ± 2)nm.

Electrostatic forces. Even if the surfaces are as clean as possible, there always remain electrostatic forces be-tween them. First, an electrostatic potential dierence

Vc still exists between clean, grounded, metallic surfaces

owing to dierences between the work functions of the

materials used [28]. Second, electrostatic forces can

re-main due to the presence of trapped charges. In liquids, these trapped charges induce double-layer forces, due to the rearrangement of ions in solution, screening the elec-trostatic interactions.

When d << R, the expression of the electrostatic force is [29]: Fe= − π0dR λD Vc2 exp(−d/λD) (6) The term Vc2

d is the contribution of the contact

poten-tial Vc between the surfaces and the term exp(−d/λD)

represents the double-layer force, screened over a distance

λD (the Debye length)[30].

As there is no free charge in nitrogen, the Debye length is innite and the electrostatic interaction is not screened, consequently there are no double-layer forces. In

ni-trogen, the contact potential was calibrated to Vc =

(87 ± 2)mV, and was compensated by an applied voltage

dierence during the measurement of the Casimir force. In contrast, in ethanol, the contact potential is strongly screened by the ions constituting the Debye layer. More-over, applying an electrostatic potential in a polar

liq-uid can yield a transcient current [31] and consequently,

charges accumulate on surfaces, preventing us from

ap-plying the method suggested in [32] to cancel the Debye

force. Therefore, we simply subtracted the contribution of electrostatic forces from force measurements, after

de-termining λD = (46 ± 6)nm and Vc = (63 ± 13)mV

[20]. In ethanol Vc is lowered because the dissociation

of molecules at the surface leads to the formation of a rst very thin screening layer of a few nm.

Measurement of the Casimir force in nitrogen. Static measurements of the Casimir force were carried out in a nitrogen atmosphere, between a gold-coated sphere and a gold-coated plate. We subtracted the vertical thermal drift by tting linearly each force curve measurement be-tween 300nm and 1µm and subtracting it from each mea-sured curve. All force curves were shifted in distance

cor-responding to the separation upon contact d0= 31nm.

The measured Casimir force is shown on Fig.2(a) for

separations ranging from 90nm to 370nm, averaging 28 independent measurements. For theoretical calculations, thermal corrections are negligible as the thermal energy

(5)

4 100 200 300 400 500 600

z [nm]

327.0 327.5 328.0 328.5 329.0 329.5

F

tot

F

Hydr o

[n

N]

(a)

1 measurement linear fit (thermal drift)

50 100 150 200 250 300

z [nm]

200 0 200 400 600 800 1000 1200 1400 1600

F

tot

F

Hydr o

F

drift

[p

N]

(b)

1 measurement fit Vc exp( z/ D) avg. 8 meas. 50 100 150 200 250 300

z [nm]

1400 1200 1000 800 600 400 200 0 200

Casimir Force [pN]

(c)

FIG. 3. Measurement of the Casimir force in ethanol. (a) Single force measurement (blue points) after subtracting the hydrodynamic force. The distance has been shifted by d0= 31nm. The red line corresponds to the linear t of the thermal drift

in the range 300nm - 600nm. (b) Single force measurement (red circles) and mean force measurement (blue points) (averaged over 8 measurements) of Fcas+ FDebye obtained after subtraction of the thermal drift and the hydrodynamic force. The mean measured force is tted by an exponential in the range d > 120nm. In this distance range, double-layer forces remove largely the Casimir force. The blue line corresponds to the exponential t, yielding the Debye length λD= (46 ± 6)nmand the surface

potential ψ0adj = (63 ± 13)mVthrough Eq. (6). (c) Measurement of the Casimir force between two Au-surfaces in a ethanol,

where blue cross correspond to the mean measured force. The orange solid line corresponds to the Lifshitz theory where the dielectric function  is evaluated from measured optical data of a real gold lm [25]; for the green dashed-dotted line,  is evaluated using the handbook optical data [26].

kBT is too small to populate the mode of lowest energy

~c/λT, as the separation distance d satises

d < 370nm  λT = ~c

kBT

≈ 7µm , at 300 K . (7)

We compared our experimental result with theoretical predictions of the Casimir force, based on optical prop-erties of Au taken from: 1) handbook of tabulated data

(green dashed-dotted line)[33], and 2) measurements on

Au samples presenting the same roughness and

prepara-tion condiprepara-tions as ours (orange solid line)[25].

The deviation from Lifshitz theory based on dielectric properties of real samples is less than 5 pN at closest sep-arations, while it reaches 10 pN at closest separations for calculations based on data from handbook. As the signal to noise ratio degrades with increasing distance, the devi-ation of the measurements from Lifshitz theory increases at larger distances. However, they remain compatible within the error bars, including systematic and statisti-cal errors (dominant at large distances). This demon-strates that surfaces must be carefully characterized for high precision measurements of the Casimir force. In or-der to make this argument more quantitative, we present

the dierence (Fexp − Fth)/Fth on Fig. 2(b), showing

the dierences between the theoretical and experimen-tal Casimir forces, for calculations based on data from handbook and calculations based on dielectric properties measured on lms with the same morphology as our lms. Measurement of the Casimir force in ethanol. In a liq-uid, the scenario is richer than in a gas because of the presence of additional eects, namely the hydrodynamic force and the Debye screening of the electrostatic inter-actions.

Measurements in ethanol were performed with the same apparatus, immediatly after the measurement in ni-trogen, so that the contact are be the same, as explained previously. During the measurement of the Casimir force, the approach velocity was chosen in order to compromise

between the hydrodynamic force FH we wanted to

mini-mize and vertical drift, limiting the time of measurement. The results presented in this paper were obtained with

100nm s−1.

In order to average the data collected from consec-utive runs, 8 data sets were acquired. First, all force curves were shifted in distance corresponding to the

sep-aration distance upon contact d0 = 31nm. Second, the

hydrodynamic force FH was subtracted from each

mea-surement independently. As the FH dependence on

dis-tance is accurately known [20], it can be safely subtracted

from the measured force. Third, to remove vertical ther-mal drift from force measurement, each force curve was

tted linearly between 300nm and 600nm. Fig. 3 (a)

shows a single force measurement in ethanol after sub-tracting the hydrodynamic force. Blue points correspond to the raw measurement, and the red line corresponds to the linear t of the thermal drift, which is then

sub-tracted from each measurement. In Fig.3 (b), we

rep-resent both a single force measurement (blue) and the average measured force (red), after subtracting the ther-mal drift and the hydrodynamic force. The remaining force shows the presence of repulsive forces at separa-tion distances larger than 60nm. These repulsive forces are attributed to the presence of ions in solution and on metallic surfaces. Then, the remaining curve

(6)

and the mean curve was tted by an exponential

func-tion A exp(−d/λD) in the distance range d > 120nm.

A and λD are adjustable parameters. In this distance

range, we assume that the Casimir force is negligible in comparison to the double-layer force (as predictions of the Lifshitz theory indicate). We obtained a Debye

length λD = (46 ± 6)nm consistent with measurements

reported by [34] and [35]. The exponential t is also

used to determine the electrostatic potential at the gold

surface ψ0 from the expression of the double-layer force

in a sphere-plane geometry: F = 4π0Rψ02/λD e−d/λD.

We evaluated the surface potential ψ0adj = (63 ± 13)mV.

After subtracting the measured double-layer force from each force measurement, the mesured Casimir force is obtained.

The measured Casimir force is presented on Fig.3(b).

The experimental data are compared to Lifshitz's theory for a gold sphere of radius R = 75µm and a gold plate separated from a distance d in ethanol. Finally the dier-ences between the theoretical predictions and the

mea-sured data are plotted in Fig.3(c). In spite of the rather

large error bars we can distinguish the two theoretical predictions: Casimir force measurements are in better agreement with Lifshitz theory based on optical proper-ties of real Au lms, presenting the same morphology as ours.

Conclusion. In conclusion, we have presented mea-surements of the Casimir force performed both in gas (nitrogen) and liquid (ethanol) environments with the same apparatus and on the same spot of the sample. The force measurements yield experimental evidence of the importance of electrostatic eects in ethanol. These eects were properly measured and subtracted, in order to determine the genuine Casimir force. Furthermore, these measurements demonstrate that the Casimir force is sensitive to changes in the optical properties of gold at distances of less than 200nm mostly in the gas en-viroment where the force is the strongest. Notably, to the best of our knowledge, this is the rst time that this inuence was measured experimentally at this range of separations. Our measurements are of signicant interest given the fundamental implications of the Casimir force in the search of new hypothetical forces, technology appli-cations of Casimir forces for micro/nano device actuation

[1,36], and the very timely nature of our measurements

[37] .

APPENDIX

This work has been supported by the ERC contract Outeucoop. We thank Irénée Frérot for helpful discus-sions.

Present address of A.Le Cunuder :

Department of Condensed Matter Physics, University of Barcelona,

Marti i Franques, 1, 080028 Barcelona, Spain

ROUGHNESS ANALYSIS

Because the Casimir force is sensitive to optical

prop-erties of gold lms [25], we characterized carefully the

topography of the surfaces. Indeed, it is commonly ac-cepted that these properties can be taken from the hand-books tabulated data. In fact, optical properties of de-posited lms depend on the method of preparation. An

interesting study [25] reported a signicant variation of

5 − 15%in Casimir force calculations, due to changes in

optical properties of Au lms.

After coating, the surface morphology of both the sphere and the plate was determined using a commer-cial AFM (Bruker). It is important to stress that these analysis were performed directly on the surfaces used in the measurement of the Casimir forces. The AFM

im-ages of a (1 × 1 µm2)sample of both surfaces are shown

in g.4. The roughness probability distribution of both

surfaces are well approximated by a Gaussian. In g.5a)

we plot the probability distribution of the sphere whose

rms roughness is wsph= (11.8 ± 0.8)nm, where the error

takes into account the AFM accuracy and the statisti-cal error based on a correlation length of about 50nm,

i.e. 400 statistical independent points on the (1 × 1 µm2)

measured surface.! " y b) x !" ! " Texxte a) Teyxte

FIG. 4. a - AFM image of 1 × 1 µm2 sample of the sphere

surface after the deposition of a 100nm thick gold lm. b AFM image 1 × 1 µm of the surface of the glass plate with a 100nmthick gold lm

The correlation length, estimated from the

Height-height correlation function plotted in g.5 b) is about

50nm. The rms roughness of the plate is wp =

(1.3 ± 0.2)nm.

This morphology analysis is then used in order to com-pare our force measurements with computations where optical properties are taken from real lms, with a simi-lar topography.

From morphology analysis, we also evaluated

approxi-matively the separation upon contact d0,rough. However,

(7)

6 101 100 101 102 r 102 101 100 101 102 H H r2a a = 0.94 ± 0.01

FIG. 5. a) Height distribution of the sphere roughness b) Height-height correlation function of the sphere surface in log-log scale. The height-height correlation function is dened as: H(r) =< [h(r) − h(0)]2 >, where h(r) is the surface height. The roughness exponent α = 0.940 ± 0.001 is extracted from the slope of the linear t and the correlation lenght ξ = 50nm is determined from the intersection between the linear t and the saturation line.

precisely d0. This analysis just helps us to check that we

nd a separation distance upon contact d0,rough of the

same order of magnitude as d0 obtained from

hydrody-namic calibration. The AFM images4 indicate that the

sphere is much rougher than the plate. Consequently, the separation distance upon contact can be evaluated

as: d0,rough = d0,sph+ wp, where wp is the rms

rough-ness of the plate and d0,sph is the highest peak of the

sphere within the contact area [38]. One estimates that

on a surface of 1µm2, there is in average only one

as-perity larger than 2.8wsph and less than one with height

larger than 3wsph, which is statistically coherent with

the image of g.4a) and the tails of the distribution in

g.5. Thus, in a contact area of about 1 µm2 one

ex-pects to nd d0,sph = 2.8wsph = (33.6 ± 2.4)nm. Notice

that this value is statistically signicant because the area involved in the force measurements are of the order of

1µm2) at d < 100nm (see ref.[24]). For the plate, the

rms roughness is wp = (1.3 ± 0.2)nm. Thus from the

topography analysis we evaluate that the maximum is

d0 < (34.9 ± 2.4)nm on an area of 1 µm2). This value

is, within error bars, statistically coherent with the hy-drodynamic calibration discussed in section 2. It is im-portant to stress that the hydrodynamic calibration is performed directly on the surfaces used in the measure-ment of the Casimir forces. Indeed, calibration were per-formed in ethanol immediatly after measurements of the Casimir force. As the liquid was introduced very car-refully in the cell after the measurement in nitrogen and as the horizontal drift of the sample is negligible, the contact are is the same during both measurements and calibration. This assumption is further justied a pos-teriori: our experimental curves all superimpose on top of each others and on top of the theoretical curves after

shifting the distance by the same value of d0. In contrast

the topography study is done on surfaces with the same statistical properties but not on the same position of the

contact area used in the experiment. Thus we use for do

the value obtained from the hydrodynamic calibration,

i.e. d=(31 ± 2)nm.

HYDRODYNAMIC CALIBRATION

The measured hydrodynamic force is plotted as a

func-tion of d in g.6 a) where it is compared to the

the-oretical force expressed in eq.(4) of main text. In the gure the measured values have been shifted

horizon-tally by do = 31nm which corresponds to the

separa-tion distance upon contact. In g.6 b), the plot of the

200 400 600 800 1000 1200 1400 0 500 1000 1500 2000 2500 3000 3500 4000 4500 distance [nm] Hy drodynamic force [ pN] 0 500 1000 1500 0 0.5 1 1.5 2 2.5 3 3.5x 10 −3 distance [nm] 1/Force [N -1]

FIG. 6. a) Hydrodynamic force in ethanol measured at v = 4.742µm s−1.b) Inverse of the hydrodynamic force as a function of d.

inverse of the force as a function of d conrms that

the value of do is good since the curve crosses the

ori-gin. The slope m of this curve is in good agreement

with the theoretical value mthdetermined from eq.(4) of

main text. Specically we nd 1/m = 1.98 10−15Jand

1/mth= 6πηR2v = 1.84 10−15J, which are in agreement

within the error bars of do, of R and of η . Thus, from

hydrodynamic calibration, we get do= (31 ± 2)nm.

CALIBRATION OF THE CONTACT POTENTIAL DIFFERENCE

An electrostatic potential dierence Vc exists between

the sphere and the plate, even if the surfaces are coated with gold and they are both electrically grounded. In-deed, there can exist a large potential dierence between clean, grounded, metallic surfaces owing to dierences between the work functions of the materials used and the

cables used to ground the metal surfaces [28]. A small

potential dierence around ten mV is sucient to over-helm the Casimir force so the contact potential dierence has to be measured and the experiment has to be carried out with a compensating voltage present at all times.

Following a procedure described in [39], we measure

the contact potential dierence Vc between the sphere

and the plate by applying an oscillating potential V =

(8)

0 200 400 600 800 1000 −100 −95 −90 −85 −80 −75 Distance [nm]

Vc

[mV]

FIG. 7. Value of the potential V2which minimizes the

excita-tion at frequency ω1. We measure a constant contact potential

dierence between 110nm et 1.1µm.

When d << R, the expression of the electrostatic force

induced by the voltage potential dierence V and Vccan

be approximated by: Fe= − π0dR d (V1cos(ω1t) + V2− Vc) 2 = −π0dR 2d [V1cos(2ω1t) + + 4V1(V2− Vc) cos(ω1t) + 2(V2− Vc)2+ V12] (8)

Because of the existence of a contact potential dierence

Vc, the system oscillates both at 2ω1 and at ω1. We

determine Vc adding a constant potential V2 until the

excitation at the frequency ω1 disappears. Indeed, when

V2= Vc, the system no longer oscillates at the frequency

ω1 (see eq.8) .

We measure the contact potential Vc as a function of

d between 1µm and 110nm. In practice, we move the

plate towards the sphere by discrete displacements. At

each separation distance dn, we measure the potential V2

which minimizes the amplitude of the oscillation at ω1.

The result of the measurement is plotted in g.7, where

we see that in our experiment Vc is constant as it is

the-oretically expected.

[1] Cyriaque Genet, Francesco Intravaia, Astrid Lambrecht, and Serge Reynaud. Electromagnetic vacuum uctua-tions, casimir and van der waals forces. 03 2003. [2] H. B. G. Casimir. On the Attraction Between Two

Perfectly Conducting Plates. Indag. Math., 10:261 263, 1948. [Kon. Ned. Akad. Wetensch. Proc.100N3-4,61(1997)].

[3] E. M. Lifshitz. The theory of molecular attractive forces between solids. Sov. Phys. JETP, 2:7383, 1956.

[4] S. K. Lamoreaux. Demonstration of the casimir force in the 0.6 to 6µm range. Phys. Rev. Lett., 78:58, Jan 1997. [5] Davide Iannuzzi, Mariangela Lisanti, and Federico Ca-passo. Eect of hydrogen-switchable mirrors on the casimir force. Proceedings of the National Academy of Sciences, 101(12):40194023, 2004.

[6] F. Chen, G. L. Klimchitskaya, V. M. Mostepanenko, and U. Mohideen. Demonstration of optically modulated dis-persion forces. Opt. Express, 15(8):48234829, Apr 2007. [7] C.-C. Chang, A. A. Banishev, G. L. Klimchitskaya, V. M. Mostepanenko, and U. Mohideen. Reduction of the casimir force from indium tin oxide lm by uv treatment. Phys. Rev. Lett., 107:090403, Aug 2011.

[8] S. de Man, K. Heeck, R. J. Wijngaarden, and D. Iannuzzi. Halving the casimir force with conductive oxides. Phys. Rev. Lett., 103:040402, Jul 2009.

[9] G. Torricelli, P. J. van Zwol, O. Shpak, C. Binns, G. Palasantzas, B. J. Kooi, V. B. Svetovoy, and M. Wut-tig. Switching casimir forces with phase-change materi-als. Phys. Rev. A, 82:010101, Jul 2010.

[10] V. B. Svetovoy, P. J. van Zwol, G. Palasantzas, and J. Th. M. De Hosson. Optical properties of gold lms and the casimir force. Phys. Rev. B, 77:035439, Jan 2008. [11] A. A. Banishev, C.-C. Chang, R. Castillo-Garza, G. L.

Klimchitskaya, V. M. Mostepanenko, and U. Mohideen. Modifying the casimir force between indium tin oxide lm and au sphere. Phys. Rev. B, 85:045436, Jan 2012. [12] R. S. Decca, D. López, E. Fischbach, G. L.

Klimchit-skaya, D. E. Krause, and V. M. Mostepanenko. Tests of new physics from precise measurements of the casimir pressure between two gold-coated plates. Phys. Rev. D, 75:077101, Apr 2007.

[13] V.B. Svetovoy and G. Palasantzas. Inuence of surface roughness on dispersion forces. Advances in Colloid and Interface Science, 216:1  19, 2015.

[14] W. Broer, G. Palasantzas, J. Knoester, and V. B. Sve-tovoy. Roughness correction to the casimir force at short separations: Contact distance and extreme value statis-tics. Phys. Rev. B, 85:155410, 2012.

[15] Ludovic Bellon. Exploring nano-mechanics through ther-mal uctuations. Accreditation to supervise research, Ecole normale supérieure de lyon - ENS LYON, Novem-ber 2010. 176 pages format A5.

[16] Jeremy N Munday, Federico Capasso, and V Adrian Parsegian. Measured long-range repulsive casimirlifshitz forces. Nature, 457(7226):170173, 2009.

[17] J. N. Munday and Federico Capasso. Precision measure-ment of the casimir-lifshitz force in a uid. Phys. Rev. A, 75:060102, Jun 2007.

[18] B. Geyer, G. L. Klimchitskaya, U. Mohideen, and V. M. Mostepanenko. Comment on precision measurement of the casimir-lifshitz force in a uid. Phys. Rev. A, 77:036102, Mar 2008.

[19] J. N. Munday and Federico Capasso. Reply to comment on `precision measurement of the casimir-lifshitz force in a uid' . Phys. Rev. A, 77:036103, Mar 2008.

[20] See Supplementary Material.

[21] Mariangela Lisanti, Davide Iannuzzi, and Federico Ca-passo. Observation of the skin-depth eect on the casimir force between metallic surfaces. Proceedings of the Na-tional Academy of Sciences, 102(34):1198911992, 2005. [22] C. C. Speake and C. Trenkel. Forces between

conduct-ing surfaces due to spatial variations of surface potential. Phys. Rev. Lett., 90:160403, Apr 2003.

(9)

8

[23] Howard Brenner. The slow motion of a sphere through a viscous uid towards a plane surface. Chemical Engi-neering Science, 16(3–4):242  251, 1961.

[24] P. J. van Zwol, V. B. Svetovoy, G. Palasantzas, and J. Th. M. De. Distance upon contact: Determination from roughness prole. Phys. Rev. B, 80:235401, 2009. [25] V. B. Svetovoy, P. J. van Zwol, G. Palasantzas, and

J. Th. M. De Hosson. Optical properties of gold lms and the casimir force. Phys. Rev. B, 77:035439, Jan 2008. [26] Edward D Palik. Handbook of optical constants of solids,

volume 3. Academic press, 1998.

[27] René I. P. Sedmik, Armando F. Borghesani, Kier Heeck, and Davide Iannuzzi. Hydrodynamic force measurements under precisely controlled conditions: Correlation of slip parameters with the mean free path. Physics of Fluids, 25(4):042103, 2013.

[28] CC Speake and C Trenkel. Forces between conducting surfaces due to spatial variations of surface potential. Physical review letters, 90(16):160403, 2003.

[29] Justine Laurent. Casimir force measurements at low temperature. Theses, Université de Grenoble, December 2010.

[30] Hans Jurgen Butt and Michael Kappl. Surface and in-terfacial forces. 08 2010.

[31] Tomlinson Fort and Robert L. Wells. Measurement of contact potential dierence between metals in liquid en-vironments. Surface Science, 12(1):46  52, 1968. [32] D. S. Ether, F. S. S. Rosa, D. M. Tibaduiza, L. B. Pires,

R. S. Decca, and P. A. Maia Neto. Double-layer force suppression between charged microspheres. Phys. Rev. E, 97:022611, Feb 2018.

[33] Edward D. Palik. Handbook of Optical Constants of Solids. Academic Press, 1997.

[34] PJ Van Zwol, G Palasantzas, and J Th M DeHos-son. Weak dispersive forces between glass and gold macroscopic surfaces in alcohols. Physical Review E, 79(4):041605, 2009.

[35] J. N. Munday, Federico Capasso, V. Adrian Parsegian, and Sergey M. Bezrukov. Measurements of the casimir-lifshitz force in uids: The eect of electrostatic forces and debye screening. Phys. Rev. A, 78:032109, Sep 2008. [36] V. B. Svetovoy and M. V. Lokhanin. Do the precise measurements of the casimir force agree with the expec-tations? Modern Physics Letters A, 15(15):10131021, 2000.

[37] Joseph L. Garrett, David A. T. Somers, and Jeremy N. Munday. Measurement of the casimir force between two spheres. Phys. Rev. Lett., 120:040401, Jan 2018. [38] M. Sedighi, V. B. Svetovoy, and G. Palasantzas. Casimir

force measurements from silicon carbide surfaces. Phys. Rev. B, 93:085434, Feb 2016.

[39] Sven de Man, Kier Heeck, and Davide Iannuzzi. Halving the casimir force with conductive oxides: Experimental details. Physical Review A, 82(6):062512, 2010.

Figure

FIG. 1. Experimental setup. A gold-coated polystyrene bead is glued on the cantilever tip which measures the  sphere-plane interaction force at distance d
FIG. 2. Measurement of the Casimir force between two Au-surfaces in a nitrogen atmosphere
FIG. 3. Measurement of the Casimir force in ethanol. (a) Single force measurement (blue points) after subtracting the hydrodynamic force
FIG. 4. a - AFM image of 1 × 1 µ m 2 sample of the sphere surface after the deposition of a 100nm thick gold lm
+3

Références

Documents relatifs

During the dry season people and animals concentrate around the few permanent water holes, and contact was made with people at such points, to explain disease control and

The immediate aim of this project is to examine across cultures and religions the facets comprising the spirituality, religiousness and personal beliefs domain of quality of life

Consider an infinite sequence of equal mass m indexed by n in Z (each mass representing an atom)... Conclude that E(t) ≤ Ce −γt E(0) for any solution y(x, t) of the damped

These criteria assume that an organization which meets this profile will continue to grow and that assigning a Class B network number to them will permit network growth

The IR or the registry to whom the IR has delegated the registration function will determine the number of Class C network numbers to assign to a network subscriber based on

Next, to con fi rm that the deoxyribose phosphate aldolase activity measured in untransfected cells was contributed by endogenous DERA, we monitored deoxyribose phosphate

Given a Hilbert space H characterise those pairs of bounded normal operators A and B on H such the operator AB is normal as well.. If H is finite dimensional this problem was solved

Abstract—This article presents a detailed description of an algorithm for the automatic detection of the mid-sagittal plane in three-dimensional (3–D) brain images.. The algorithm