arxiv:1311.3265v2 [cond-mat.supr-con] 15 nov 2013

12
arXiv:1311.3265v2 [cond-mat.supr-con] 15 Nov 2013 Noname manuscript No. (will be inserted by the editor) A. Casey · F. Arnold · L. V. Levitin · C. P. Lusher · J. Saunders · A. Shibahara · H. van der Vliet · D. Drung · Th. Schurig · G. Batey · M. N. Cuthbert · A. J. Matthews Current Sensing Noise Thermometry: A fast practical solution to low temperature measurement Received: date / Accepted: date Abstract We describe the design and performance of a series of fast, precise current sensing noise thermometers. The thermometers have been fabricated with a range of resistances from 1.290 Ω down to 0.2 mΩ. This results in either a thermometer that has been optimised for speed, taking advantage of the improvements in superconducting quantum interference device (SQUID) noise and bandwidth, or a thermometer optimised for ultra-low temperature measurement, minimising the system noise temperature. With a single tem- perature calibration point, we show that noise thermometers can be used for accurate measurements over a wide range of temperatures below 4 K. Comparisons with a melting curve thermometer, a calibrated germanium thermometer and a pulsed platinum nuclear magnetic resonance thermome- ter are presented. For the 1.290 Ω resistance we measure a 1 % precision in just 100 ms, and have shown this to be independent of temperature. Keywords Thermometry · SQUID A. Casey, F. Arnold, L. V. Levitin, C.P. Lusher, J. Saunders, A. Shibahara, H. van der Vliet Department of Physics, Royal Holloway University of London, Egham, Surrey, TW20 0EX, United Kingdom Tel.: +44 (0)1784 414351 Fax: +44 (0)1784 472794 E-mail: [email protected] D. Drung, Th. Schurig Physikalisch-Technische Bundesanstalt, Abbestrasse 2-12, D-10587 Berlin, Ger- many G. Batey, M. N. Cuthbert, A. J. Matthews Oxford Instruments Omicron Nanoscience, Tubney Woods, Oxon, OX13 5QX, United Kingdom E-mail: [email protected]

Upload: others

Post on 24-Dec-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: arXiv:1311.3265v2 [cond-mat.supr-con] 15 Nov 2013

arX

iv:1

311.

3265

v2 [

cond

-mat

.sup

r-co

n] 1

5 N

ov 2

013

Noname manuscript No.(will be inserted by the editor)

A. Casey · F. Arnold · L. V. Levitin ·

C. P. Lusher · J. Saunders · A.

Shibahara · H. van der Vliet · D.

Drung · Th. Schurig · G. Batey · M.

N. Cuthbert · A. J. Matthews

Current Sensing Noise Thermometry:A fast practical solution to lowtemperature measurement

Received: date / Accepted: date

Abstract We describe the design and performance of a series of fast, precisecurrent sensing noise thermometers. The thermometers have been fabricatedwith a range of resistances from 1.290 Ω down to 0.2 mΩ. This results ineither a thermometer that has been optimised for speed, taking advantage ofthe improvements in superconducting quantum interference device (SQUID)noise and bandwidth, or a thermometer optimised for ultra-low temperaturemeasurement, minimising the system noise temperature. With a single tem-perature calibration point, we show that noise thermometers can be usedfor accurate measurements over a wide range of temperatures below 4 K.Comparisons with a melting curve thermometer, a calibrated germaniumthermometer and a pulsed platinum nuclear magnetic resonance thermome-ter are presented. For the 1.290 Ω resistance we measure a 1 % precision injust 100 ms, and have shown this to be independent of temperature.

Keywords Thermometry · SQUID

A. Casey, F. Arnold, L. V. Levitin, C.P. Lusher, J. Saunders, A. Shibahara, H. vander VlietDepartment of Physics, Royal Holloway University of London, Egham, Surrey,TW20 0EX, United KingdomTel.: +44 (0)1784 414351Fax: +44 (0)1784 472794E-mail: [email protected]

D. Drung, Th. SchurigPhysikalisch-Technische Bundesanstalt, Abbestrasse 2-12, D-10587 Berlin, Ger-many

G. Batey, M. N. Cuthbert, A. J. MatthewsOxford Instruments Omicron Nanoscience, Tubney Woods, Oxon, OX13 5QX,United KingdomE-mail: [email protected]

Page 2: arXiv:1311.3265v2 [cond-mat.supr-con] 15 Nov 2013

2

1 Introduction

There is a rapidly expanding use of low temperature platforms (T < 1 K), forfundamental science, sensitive instrumentation, and new technologies of po-tentially significant commercial impact. The availability of commercial “pushbutton” cryogen-free dilution refrigerators is a key game-changing recent de-velopment, in which European companies have played the leading role. Theavailability of cryogen-free commercial platforms to 1 mK is a future prospect[1].

Precise measurement of the thermodynamic temperature of these lowtemperature platforms is crucial for their operation. However, despite thesophisticated nature of the cooling technology, reliable and convenient ther-mometers are unavailable.

In the temperature range below 1 K the dissemination of the kelvin isusually realised using secondary thermometers and superconducting refer-ence points [2,3,4]. Examples include: resistance thermometers; paramag-netic susceptibility; vibrating wire/tuning fork viscometry. These thermome-ters do not reliably cover the temperature range 1 K to 10 mK (a fortiorito 1 mK). The calibration of these thermometers against the PLTS-2000 [5,6] is impractically time consuming and expensive. Resistance thermometersare intrinsically subject to shifts in calibration and prone to self heatingand pick-up of electromagnetic interference. Paramagnetic-salt susceptibilitythermometers are subject to degradation with time, and viscometric meth-ods measure temperature inside the mixing chamber and not on the typicalsample mounting platform.

The dependence of 3He melting pressure on temperature has been usedas the laboratory standard in specialist laboratories for many years. ThePLTS-2000 scale is based on the melting pressure of 3He but this is viewedas difficult to implement for the rapidly expanding community of users ofultra-low temperature platforms. 60Co nuclear orientation thermometry is aprimary technique that relies on the known anisotropy of the decay of ra-dioactive nuclei as a function of temperature but is only effective in a narrowtemperature range, and is slow and user unfriendly. Nuclear magnetic suscep-tibility thermometry is likewise insensitive above 20 mK, and user unfriendly.Reliable and traceable thermometry below 1 K thus offers demanding chal-lenges even to the specialist, and, despite the sophisticated low temperatureplatforms now readily available with “push-button” operation, a solutionwhich is reliable, traceable and user-friendly has yet to be developed.

A quote attributed to Prof O V Lounasmaa, founder of the Low Tempera-ture Laboratory at Helsinki University of Technology (now Aalto University)in the context of the T <1 K regime reads “A person with one thermometerknows what the temperature is; a person with two will forever have doubt”.The challenge is to rectify this state of affairs.

The need to have a thermometer that measures thermodynamic temper-ature outside of international standards laboratories has led to the devel-opment of thermometers with a simple relationship between the measuredparameter and the Boltzmann constant, kB, for example thermometers thatcouple noise to a SQUID either directly [7] or indirectly via the magnetic

Page 3: arXiv:1311.3265v2 [cond-mat.supr-con] 15 Nov 2013

3

Johnson noise [8,9,10,11] and Coulomb Blockade thermometers [12] basedon electric conductance characteristics of tunnel junction arrays.

In this paper we describe recent advances in current sensing noise ther-mometer, exploiting low noise and high bandwidth dc SQUID amplifiers, forthe high precision and rapid measurement of absolute temperature, coveringthe entire temperature range of interest from 4K to below 100 microK.

2 Current Sensing Noise Thermometry

The current sensing noise thermometer was pioneered by Giffard et al.[13,14] and developed by Webb and Washburn [15]. The principle is to measurethe thermal Johnson noise [16] in a resistor to determine its temperature.The performance of the thermometers developed in [13,14,15] was limitedby the noise of the RF SQUID amplifiers; in [15] the typical device noisetemperature was a few mK. Using a low transition temperature dc SQUIDas the front end amplifier the current sensing technique becomes a practi-cal approach for measuring absolute temperature from 4.2 K down to submK temperatures, as shown in [7]; this work used commercial dc SQUIDs.The improved sensitivity of dc SQUIDs fabricated at PTB [17] allows theuse of sensor resistors with much larger resistance, thereby increasing thespeed of the thermometer. In this paper we demonstrate experimentally animprovement in speed by orders of magnitude and we demonstrate that thepercentage precision is independent of temperature.

The principle of operation of a current sensing noise thermometer is dis-cussed in detail in our earlier paper [7] but is summarised below. The meansquare open circuit noise voltage per unit bandwidth across the terminals ofa resistor, R, at temperature T is given by the Nyquist expression:

V 2

N

= 4kBTR (1)

This is a white noise spectrum over the frequency range of interest. In ourcurrent sensing noise thermometer, the resistor whose absolute temperatureis to be measured is connected in series with the superconducting input coilof the readout SQUID through a superconducting twisted pair. The meansquare noise current flowing in the SQUID input coil per unit bandwidtharising from the thermal noise in the resistor is then given by:

I2N⟩

=4kBT

R

(

1

1 + ω2τ2

)

(2)

where ω = 2πf . The time constant τ = L/R, where the total inductancein the input circuit L includes the SQUID input coil inductance Li and anyadditional stray inductance including that of the twisted pairs. By increasingthe resistance, it is possible to increase the bandwidth of the noise spectrum.However, this results in a reduced noise power from the resistor at ωτ < 1,which must be kept much greater than the SQUID noise in order to ensurea reasonable device noise temperature, TN, much less than the minimumtemperature we wish to measure . The SQUID is operated in flux-locked loop(FLL) mode where the gain depends only on fixed, known parameters. Thus,

Page 4: arXiv:1311.3265v2 [cond-mat.supr-con] 15 Nov 2013

4

Fig. 1 Schematic diagram of the current sensing noise thermometer. The dcSQUID amplifies the thermal noise current of the resistor with the output cap-tured by a digitiser. T1 can be at a different temperature to T2, the temperature ofthe SQUID.

if the value of the resistive sensor and the SQUID gain can be measured, acurrent noise spectrum can be used to determine absolute temperature usingEquation 2.

3 Experimental Details

Figure 1 shows a schematic of the measurement set-up. The SQUID sensorused in this work is a PTB two-stage SQUID [17] with an input coil induc-tance of Li = 1.05 µH and a mutual inductance Mi between the SQUIDand the input coil equivalent to 1/Mi = 0.29 µA/φ0, where φ0 is the fluxquantum. When operated with a 10 kΩ feedback resistor, the FLL gain is0.42 V/φ0. The SQUID is located remotely from the sensor resistor, allowingthe resistor to be placed in environments hostile to SQUID operation. In thiscase the SQUID is mounted on the continuous heat-exchanger plate of thedilution fridge at approximately 100 mK. At this temperature, the white fluxnoise floor of the SQUID is 0.58 µφ0Hz

−1/2, equivalent to a coupled energysensitivity, ǫc = 22h in units of Planck’s constant, where:

ǫc =1

2Li

φ2

N

M2

i

(3)

and⟨

φ2

N

is the flux noise power. This can be compared to ǫc = 500h forthe commercial Quantum Design [18] SQUID operated at 4.2 K used in ourprevious work [7,19]. This lower noise floor corresponds to a lower amplifiernoise temperature TN. Thus through choice of resistor the user can min-imise the noise temperature for absolute precision or increase the bandwidthof the measurement for increased speed. The primary focus of this workwas to establish the optimum choice of resistance value for operation of thethermometers on a dilution refrigerator (with a typical base temperature of5 mK). In addition in order to investigate the 3He melting curve fixed pointson the provisional low temperature scale PLTS-2000 we need a thermometerthat can be reliably operated below 1 mK. For these reason we chose sensorresistance values of 0.2 mΩ, 0.14 Ω and 1.290 Ω.

Page 5: arXiv:1311.3265v2 [cond-mat.supr-con] 15 Nov 2013

5

Fig. 2 Schematic of a PtW noise thermometer mounted onto a copper sampleplate. The plate has either a thermal link to the mixing chamber of the dilutionfridge or the sensor is mounted directly onto the copper nuclear demagnetisationstage.

3.1 Construction of noise sensors

The 0.2 mΩ resistor was fabricated with a copper foil as described in [7].Using this material it is not practical to make a sensor with a resistancegreater than ≈ 10 mΩ. A higher resistivity material was needed for the 0.14 Ωand 1.290 Ω sensors. For this work 50 µm diameter platinum tungsten wire(Pt92%/W8%) [20] was used. PtW was chosen as it has a weakly temperaturedependant resistivity as well as having a very small magneto-resistance [21].Figure 2 is a schematic diagram of the sensor resistor and the sample plate.The required length of PtW wire is achieved by hard soldering it to two piecesof 100 µm thick copper foil separated by a well-defined gap, using JohnsonMatthey EasyFlo2 solder. One foil is electrically grounded and consequentlyheat sunk to a large annealed copper holder with a copper screw contact. Theother foil is electrically isolated from the holder. The sensor resistor is thenconnected to a pair of niobium screw terminals via short sections of CuNi cladNbTi wire. The screw terminals are connected to the remote SQUID inputcoil through a shielded NbTi superconducting twisted pair. The resistor andcopper former are enclosed in a niobium can for shielding and heat sunk to acopper sample plate with an M4 screw and cone joint as depicted in Figure2. For the tests of the 1.290 Ω sensor the sample plate was mounted at themixing chamber plate of the dilution fridge, the 0.2 mΩ and 0.14 Ω sensorswere mounted directly on the top plate of a copper nuclear demagnetisationstage.

3.2 Characterisation of the sensors

In principle, the current sensing noise thermometer is a primary thermome-ter; if the resistance of the sensor and the transfer function of the SQUID

Page 6: arXiv:1311.3265v2 [cond-mat.supr-con] 15 Nov 2013

6

Fig. 3 Current noise spectra at 500 mK for the three noise sensors evaluated. Solidlines are fits using equation 2.

are known accurately. In practice however, it is more convenient to stabiliseat a known reference temperature and determine the parameters R and Lfrom a fit of the output noise power. For this single point calibration of ourthermometers, the sample plate was held at 500 mK, measured by the melt-ing curve thermometer. The temperature stabilisation was realised with adigital PID temperature controller to within 5 µK. The SQUID output volt-age is read out using room temperature Magnicon XXF electronics [22] andis captured with a PXI digitiser card [23]. Time traces were captured withsampling frequencies of 50 kS/s, 500 kS/s, 2.5 MS/s for 0.2 mΩ, 0.14 Ω and1.290 Ω sensors respectively, with a record length of 220 points. 1000 timedomain traces were captured for each sensor. Each captured time trace is in-dividually Fourier transformed and the power spectral density is averaged inthe frequency domain for improved precision. Before fitting, it was necessaryto remove the noise power contribution of the SQUID, which was measuredat high frequencies and assumed to be frequency independent over the rangeused for the fits. The current noise spectra obtained in this way are shown infigure 3. Fits were performed on the resulting current noise spectrum usingEquation 2, with T = 500 mK. The resulting values for the resistance of thenoise sensor R and total inductance of the circuit L are summarised in table1. These parameters were constrained in the fitting routine to determine thetemperature of subsequent noise traces.

An alternative method to obtain the single fixed point calibration is tointegrate a superconducting fixed point directly into the input circuit of thethermometer. An additional coil can be added to the superconducting fluxtransformer,wound around a sample with a known superconducting transi-tion temperature; thermalized at the same point as the resistive sensor. Thetotal inductance, L, of the input circuit is then sensitive to the state of thesuperconducting fixed point. Figure 4 shows the temperature dependence ofthe inductance measured in a current sensing noise thermometer in which we

Page 7: arXiv:1311.3265v2 [cond-mat.supr-con] 15 Nov 2013

7

Fig. 4 Total inductance, L, of the noise thermometer input circuit as a functionof temperature measured on the calibrated Germanium resistance thermometer. Asharp decrease in inductance is observed when the noise thermometer cools throughthe superconducting transition of the integrated tantalum fixed point device.

Table 1 Summary of the parameters of the noise thermometers evaluated, TN ismeasured from the high frequency noise.

Resistance Total Inductance TN Bandwidth (3dB)

0.20436 mΩ 1.5016 µH <1 µK 22 Hz0.13781 Ω 1.5483 µH 40 µK 14.5 kHz1.290 Ω 1.315 µH 400 µK 156 kHz

have included an in-situ tantalum [24] fixed point device ( T c = 4.45 K). Here12 turns of NbTi wire are wound around a tantalum bobbin forming part ofthe superconducting flux transformer. The sharp transition of the tantalumis clearly visible in the noise thermometer measurement.

4 Performance

The current sensing noise thermometers were operated over a wide tempera-ture range from 4.2 K down to 200 µK. Over this range there is not a singlethermometer against which a comparison can be made. In this work we useda 3He melting curve thermometer to determine the Provisional Low Temper-ature Scale, PLTS-2000 (between 1 and 700 mK); a calibrated germaniumthermometer from 50 mK to 4.2 K and a pulsed NMR platinum thermometer,calibrated by the 3He melting curve thermometer, was used below 1 mK. Fig-ure 5 shows this comparison for all three thermometers. At each temperatureabove 6 mK the temperature was stabilised for 1 hour. Excellent agreementbetween the thermometers was observed over a wide temperature range. De-

Page 8: arXiv:1311.3265v2 [cond-mat.supr-con] 15 Nov 2013

8

0.1 1 10 100 10000.1

1

10

100

1000

Noi

se th

erm

omet

er te

mpe

ratu

re [m

K]

T (MCT/Germanium/PLM) [mK]

R = 1.29 R = 0.14 R = 0.2 m

TNT = 0.999 T + 0.002 mK

Fig. 5 Comparison of temperature obtained from the current sensing noise ther-mometers with that of a 3He melting curve thermometer, a calibrated germaniumresistance thermometer and a pulsed platinum NMR thermometer over a wide tem-perature range. The linear fit (solid line) to high temperature data is also shown.The decoupling of the thermometer at the lowest temperatures is modelled by fitsthat assume a metallic thermal link to the nuclear stage with a constant heat leakinto the noise thermometer (dotted lines).

viations from the reference thermometer are observable up to progressivelyhigher temperature for higher noise sensor resistance. This can be accountedfor by a heat leak to the resistor resulting in a temperature gradient acrossthe resistor. Reducing the size of the heat leak through improving the heatsinking and filtering of the wires to the ‘floating’ side of the sensor resistorwill be the subject of future development.

Figure 6 shows the percentage deviation of the comparison between the0.14 Ω sensor and the melting curve thermometer. Here an agreement tobetter than 1 % is achieved down to the lowest temperatures attainable ona commercial dilution refrigerator.

It is clear that below 1 mK a relatively low sensor resistance is desirableto ensure good thermalization. The non-contact measurements, which formthe basis of the magnetic field fluctuation thermometer [11] are an alternativemethod for T < 100µK. At the lowest temperature obtained, 165 µK on theplatinum thermometer, the 0.2 mΩ sensor cooled to 180 µK.

Another figure of merit for a thermometer is the measurement time re-quired to obtain a given precision of temperature measurement. If we considercapturing a time trace of N points with a time ∆t between each point, foreach point to be an independent measure of the square of the noise current,∆t must be of order the correlation time of the noise or greater (which forthis system is the time constant τ = L/R). Therefore in a measuring timetmeas one can make approximately N = tmeas/τ independent measurements.A rough estimate for the precision in a given measuring time is thus given

Page 9: arXiv:1311.3265v2 [cond-mat.supr-con] 15 Nov 2013

9

5 10 100 500

0.01

0.1

1

10

|TN

T - T M

CT|/T

MC

T [%]

MCT Temperature [mK]

Fig. 6 Comparison of temperature obtained from the 0.14 Ω current sensing noisethermometers with the 3He melting curve thermometer, down to the lowest tem-peratures typically reached using a dilution refrigerator. Plotted as a percentage of∆T/T .

by:

∆T

T≈

(

2

N

)1/2

=

(

tmeas

)1/2

(4)

The percentage precision in a given measuring time is thus independent oftemperature and determined by the noise bandwidth. The noise bandwidthfor the sensors used is given in table 1, and can be seen in figure 3. Increasingthe resistance of the sensor results in an improved precision in a given time,which was the motive behind using PtW resistors with a larger resistancevalue.

To experimentally determine the precision of the 1.29 Ω PtW thermome-ter, the sample plate was stabilised at a known fixed temperature. Single shotnoise traces were taken consecutively. Approximately 1000 individual spectrawere captured for accurate statistics. A sampling frequency of 2.5 MS/s with220 points was used, giving a measurement time of 419 ms per trace. Figure7A shows data taken with the PtW thermometer stabilised at 493 mK usingthe germanium thermometer. This follows a normal distribution as shown byFig.7B, where a Gaussian fit to a histogram of temperatures is shown. Theprecision is calculated using σ/T , where σ is the standard deviation of thedata and T is the mean temperature. In a measurement time of 419 ms, aprecision of 0.50 % was obtained. This precision measurement was repeatedat 1.5 K, 100 mK and at 13 mK; the percentage precision was confirmed tobe independent of temperature. Reducing the measurement time to 105 ms,resulted in a temperature independent precision of 1.15 %. We contrast thiswith the pioneering work on current sensing noise thermometers [14], whichrequired a measurement time of 2500 s to achieve a similar performance. The

Page 10: arXiv:1311.3265v2 [cond-mat.supr-con] 15 Nov 2013

10

0 500 1000 1500 2000 2500 3000

485

490

495

500

Te

mpe

ratu

re [m

K]

Time [s]

484 486 488 490 492 494 496 498 500 5020

10

20

30

40

50

60

Cou

nts

Temperature [mK]

Histogram of TN

Gaussian fit

A

B

Fig. 7 (colour online) A: The scatter of temperature measurements using the1.29 Ω PtW current sensing noise thermometer, stabilised with a germanium ther-mometer at 493 mK. The x axis shows the real time taken to capture the traces.Each temperature reading was acquired in a measurement time of 419 ms. Thehorizontal line of best fit is the mean temperature. B: A histogram of temperaturesshowing the scatter follows a normal distribution.

1.29 Ω PtW thermometer is about two orders of magnitude faster than ourprevious noise thermometers [19], magnetic field fluctuation thermometers [9]and on-chip current sensing noise thermometers [10], which all have achieved1 % precision in times of the order of 10 s. Ultra-low temperature measure-ments using a contact free cross-correlation noise thermometer [11] have beenreported down to 45 µK, here using two SQUIDs the cross-correlation tech-nique allows them to reduce the noise contribution from SQUID amplifiers,and the reported measurement time was 400 seconds.

Figure 8 illustrates the dependence of the precision obtained from allthree of the current sensing noise thermometers as a function of the mea-surement time at a temperature of 100 mK. To vary the measurement time,the sampling frequency was kept constant at 50 kS/s, 500 kS/s, 2.5 MS/sfor 0.2 mΩ, 0.14 Ω and 1.290 Ω respectively; 220 points were captured foreach trace and the number of averages was changed to give total measure-ment times of between 0.1 and 250 seconds. The best fit lines shown areproportional to the inverse of the square root of the measurement time, asin Equation 4. The measured precision of 0.03 % obtained with the 1.29Ω,sensor in a measurement time of 100 s, is approaching the limit of our abilityto stabilise the temperature at 100 mK.

Page 11: arXiv:1311.3265v2 [cond-mat.supr-con] 15 Nov 2013

11

0.1 1 10 1000.01

0.1

1

1.29 0.14 0.2 m 1/t0.5

Time [Sec]

Pre

cisi

on [%

]

Fig. 8 Measured precision as a function of measurement time for each of the threethermometers tested, a best fit line proportional to the inverse of the square of themeasurement time is shown for the 1.29 and 0.14 Ω sensors.

5 Conclusions

In conclusion, a series of fast, precise current sensing noise thermometersbased on a range of resistive sensors has been tested. Using a single temper-ature calibration, the noise thermometers can be used to accurately measurea wide range of temperatures. We have presented data on a current sens-ing noise thermometer that has been optimised for speed, taking advantageof the improvements in SQUID noise and bandwidth. This is two orders ofmagnitude faster than previous noise thermometers, achieving 1.15 % preci-sion in just 0.1 seconds. The 0.14 Ω sensor tested is in agreement with the3He melting curve thermometer to within better than 1 % over the operat-ing temperature range of a dilution refrigerator, and in this case less than2 seconds of measurement time is required for 1 % precision. Finally thelowest value of resistance chosen, 0.2 mΩ resulted in a thermometer with anoise temperature of 1 µK that showed reasonable agreement with a pulsedplatinum NMR thermometer to below 200 µK.

The current sensing noise thermometers described here constitute prac-tical devices for disseminating thermodynamic temperature. The system isuser-friendly and could be installed on any commercial ultra-low temperatureplatform, to provide a reliable and traceable measure of thermodynamic tem-perature, with high precision (0.1 %) and high accuracy (better than 1%) ina reasonable measurement time (< 10 s).The potential provision of absolutetemperature to all users promises to be a revolutionary technical develop-ment in a temperature range expected to be of growing technological andcommercial significance.

Page 12: arXiv:1311.3265v2 [cond-mat.supr-con] 15 Nov 2013

12

Acknowledgements This work was supported by the European Metrology Re-search Program (EMRP). The EMRP is jointly funded by the EMRP participatingcountries within EURAMET and the European Union. Additional support wasthrough the MICROKELVIN project. We acknowledge the support of the Euro-pean Community - Research Infrastructures under the FP7 Capacities SpecificProgramme, MICROKELVIN project number 228464.

References

1. G. Batey, A. Casey, M. N. Cuthbert, A. J. Matthews, J. Saunders, A. Shibahara,accepted for New J. Phys. (2013).

2. O. V. Lounasmaa, Experimental principles and methods below 1 K, AcademicPress (1974).

3. C. Enss, S. Hunklinger, Low Temperature Physics, Springer (2005).4. F. Pobell, Matter an Methods at Low Temperatures, Springer (2007).5. BIPM, Proces-Verbaux des Seances du Comite International des Poids etMesures 68, 128 (2001).

6. R. L. Rusby, M. Durieux, A. L. Reesink, R. P. Hudson, G. Schuster, M. Kuhne,W. E. Fogle, R. J. Soulen, D. E. Adams, J. Low Temp. Phys. 126, 633 (2002).

7. C. P. Lusher, Junyun Li, V. A. Maidanov, M. E. Digby, H. Dyball, A. Casey, J.Nyeki, V. V. Dmitriev, B. P. Cowan and J. Saunders, Meas. Sci. Technol. 12, 1(2001).

8. T. Varpula and H. Seppa, Rev. Sci. Instrum. 64(6), 1593 (1993).9. J. Beyer, D. Drung, A. Kirste, J. Engert, A. Netsch, A. Fleischmann, and C.Enss, IEEE Trans. Appl. Supercond. 17, 2 (2007).

10. J. Engert, J. Beyer, D. Drung, A. Kirste, D. Heyer, A. Fleischmann, C. Enssand H.-J. Barthelmess, J. Phys. C 150, 12012(2009).

11. D. Rothfuß, A. Reiser, A. Fleischmann, C. Enss, Appl. Phys. Let. 103, 052605(2013).

12. J. P. Pekola, K. P. Hirvi, J. P. Kauppinen, and M. A. Paalanen, Phys. Rev.Lett. 73, 2903 (1994).

13. R. P. Giffard, R. A. Webb, J. C Wheatley, J. Low Temp. Phys. 6, 533-610(1972).

14. R. A. Webb, R. P. Giffard, J. C. Wheatley, J. Low Temp. Phys. 13, 383-429(1973).

15. R. A. Webb, S. Washburn, AIP Conference Proceedings, 103, 453-466 (1983)16. J. B. Johnson, Phys. Rev. 32, 97 (1928).17. D. Drung, C. Aßmann, J. Beyer, A. Kirste, M. Peters, F. Ruede, Th. Schurig,IEEE Trans. Appl. Supercond. 17, 699 (2007).

18. Quantum Design, 11578 Sorrento Valley Road, San Diego, CA 92121-1311,USA.

19. A. Casey, B. P. Cowan, H. Dyball, Junyun Li, C. P. Lusher, V. Maidanov, J.Nyeki, J. Saunders, Dm. Shvarts, Physica B 329-333, 1556 (2003).

20. Goodfellow Cambridge Ltd, Ermine Business Park, Huntingdon, England PE296WR.

21. G. S. Cieloszyk, P. J. Cote, G. L. Salinger, and J. C. Williams, Rev. Sci.Instrum. 46, 1182 (1975).

22. Magnicon GmbH, Lemsahler Landstr. 17, Hamburg, Germany.23. National Instruments Corporation, 1105 N Mopac Expressway, Austin,TX78759-3504, USA.

24. J. G. C. Milne, Phys. Rev. 122, 387388 (1961).