34
JPICH Heritage in CHanging Environments Call for Proposals - Final version 04 September, 2017 p. 1 The JHEP2 project has received funding from the European Union's Horizon 2020 research and innovation programme under grant agreement No 6995327 JPI Cultural Heritage and Global Change JPICH Heritage in Changing Environments Call Deadline for submission of proposals: 30 November 2017, 14:00 CET. In developing the research proposals, applicants must follow the guidelines and the application structure as explained below. Proposals should demonstrate a strong fit to the overall aims and research topics of the Joint Programming Initiative on Cultural Heritage and Global Change (JPICH) as well as to the call on cultural heritage in Changing Environments itself. 1. Introduction The JPICH ‘HERITAGE IN CHANGING ENVIRONMENTS’ Call is designed to support the development of new, research-based ideas and knowledge in response to the rapidly and widely changing context with which heritage and heritage practice is faced. It invites research projects that help cultural heritage to meet societal challenges and contribute to the development of society. The call aims to fund excellent research that is collaborative, transnational, interdisciplinary and innovative. Three broad categories of the changing environments of heritage are addressed in this call: changing (physical) environments; changing social and economic environments; and changing political and cultural environments. Projects funded through this call will use cultural heritage to address global challenges such as the impacts of climate change, environmental deterioration, migration, demographic and social change, and diasporic change, urbanisation and de-ruralisation, economic inequity, changing perceptions and sustainability. Many other important themes cross-cut these topics, notably the benefits of interdisciplinarity and the integration of different approaches both within and beyond the heritage field. An important aspect of projects within this call will be the relationships between researchers and a broad range of stakeholders outside the academic sector, and the potential to influence policy-makers in key areas of heritage management and planning. The aims of the call are: to support well-defined, innovative, interdisciplinary, internationally-focussed and collaborative research projects of the highest quality and standards that will lead to significant advances in cultural heritage practice across a broad range of research communities and societies, notably through trans-frontier exchange and in international contexts; to maximise the value of research outcomes by promoting knowledge exchange with individuals and organisations outside the immediate research community, to include policy makers, businesses and commercial enterprises, the broader heritage sector, voluntary and community groups and the general public; to support a range of interactions and partnerships between cultural heritage researchers and a variety of user communities, to include policy makers, businesses and commercial enterprises, the broader heritage sector, voluntary and community groups and the general public, thereby maximizing the value of the research outcomes.

Johnson Noise - Semantic Scholar · mal noise, the Boltzmann constant was measured to an accuracy of»4%.7 Alternatemethods implementing a undergraduate physics experiment on Johnson

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Johnson Noise - Semantic Scholar · mal noise, the Boltzmann constant was measured to an accuracy of»4%.7 Alternatemethods implementing a undergraduate physics experiment on Johnson

Johnson Noise

Nichols A. RomeroJunior Physics Laboratory, Massachusetts Institute of Technology

Cambridge, Massachusetts 02139

(Novembber 26, 1998)

An experiment was performed that determines the Boltz-mann constant k and the centigrade temperature of abso-lute zero by measuring the thermal noise of resistors. TheNyquist theorem provides a quantitative relationship be-tween the thermal electromotive force across a conductorand its resistance and temperature. Measurement of theroot-mean-square RMS voltage for a variety of resistors ata fixed temperature was used to calculate the Boltzmannconstant. The RMS voltage for a 22.5 kΩ resistor was mea-sured over 300 degree temperature range. This latter dataextrapolated to zero centigrade gave an estimate of abso-lute zero and provided an additional method for determiningthe Boltzmann constant. The experimentally determined val-ues of the Boltzmann constants, 1.37 ± 0.06 × 10−23 J/K &1.363 ± 0.025 × 10−23 J/K, and the centigrade temperatureof absolute zero, −265.5± 6.9C, are in good agreement withthe accepted values.

I. INTRODUCTION

This paper is a full report on the junior lab experi-ment: Johnson Noise. In this experiment, we studythe phenomenon of thermal (Johnson) noise as predictedby the Nyquist Theory.

This report has been partitioned into sections accord-ingly, each discussing a specific aspect of the experiment.Section II discusses the theoretical background relevantto the experiment by deriving the Nyquist Theorem usingtwo different approaches. The experimental apparatusand details of its operation are discussed in section III.Section IV presents the experimental results. Concludingremarks are given in section V.

II. NYQUIST THEORY

Johnson Noise is the mean-square electromotive forcein conductors due to thermal agitation of the electro-magnetic modes which are coupled to the thermal envi-ronment by the charge carriers. The Nyquist Theory isof great importance to experimental physics and in elec-tronics. It gives a quantitative expression for the John-son Noise generated by a system in thermal equilibriumand is therefore needed in any estimate of the limitingsignal-to-noise ratio of an experimental apparatus. Inthis section, the Nyquist theorem is derived in two ways:first, following the original transmission line derivation,and, second using microscopic arguments [1], [2].

A. Transmission Line Derivation

Consider two conductors each of resistance R at a tem-perature T connected as depicted in Figure 1. Conductor1 produces a current I in the circuit equal to the electro-motive force due to thermal agitation divided by the totalresistance 2R. This current delivers power to conductor2 equal to current squared times the resistance. By sym-metry, one can deduce that the situation is reciprocal.Conductor 2 produces a similar current which deliverspower to conductor 1. Because the two conductors are atthe same temperature, the second law of thermodynam-ics dictates that the power flowing in both directions isequal. I emphasize that no assumption about the natureof conductors has been made.

1 2R R

FIG. 1. Two conductors with equal resistance R.

It can be shown that this equilibrium condition holdsat any given frequency. Suppose there exists a frequencyinterval ∆ν1 where conductor 1 receives more power thanit transmits. We then connect a non-dissipative networkwith a resonance in the frequency interval ∆ν1 betweenthe two conductors (refer to Figure 2). Since the sys-tem was in equilibrium prior to inserting the network, itfollows that after is insertion more power would be trans-ferred from conductor 2 to conductor 1. However, as theconductors are at the same temperature, this would vio-late the second law of thermodynamics. The results wehave arrived at are important enough to merit summariz-ing. By eminently reasonable theoretical arguments, wecan conclude that the electromotive force due to thermalagitation in conductors are universal functions of (referto Figure 3):

• frequency ν

• resistance R

• temperature T

Experiments performed by Dr. J. B. Johnson in 1928confirmed the formula which was later derived Dr. H.Nyquist on purely theoretical grounds [3].

The derivation of the mean-square voltage 〈V 2〉 acrossa conductor closely follows Nyquist’s original derivation.The problem of determining a quantitative expression forthe thermal agitation (i.e. the mean-square voltage) of

1

Page 2: Johnson Noise - Semantic Scholar · mal noise, the Boltzmann constant was measured to an accuracy of»4%.7 Alternatemethods implementing a undergraduate physics experiment on Johnson

1 2R R

FIG. 2. Two conductors plus resonant circuit.

FIG. 3. Voltage-squared vs. resistance component for var-ious types of conductors.

the conductor can be viewed as a simple one-dimensionalcase of black-body radiation. Consider a lossless one-dimensional transmission line of length L terminated atboth ends by conductors with resistance R. The trans-mission line has been chosen to have a characteristicimpedance Z = R; consequently any voltage wave propa-gating along the transmission line is completely absorbedby the terminating resistor without any reflections. Volt-age waves of the form V = V0 exp [i(kxx− ωt)] propagatedown the transmission line at velocity v = ω/kx. Theavailable number of modes can be calculated by impos-ing the periodic boundary condition V (0) = V (L) on thepropagating voltage waves. The wave vector kx is relatedto the length by the relation kxL = 2πn where n is anyinteger. The density of modes is then,

D(ω) =1

L

dn

=1

L

dn

dkx

dkxdω

=1

2πv(1)

The mean energy per mode is given by the Planck for-mula,

〈ε(ω)〉 =hω

exp hωkT − 1

(2)

〈ε(ω)〉 ≈ kT (3)

where in the last line we made use of the equipartitiontheorem: in the classical limit, hω ¿ kT , each squaredcanonical term in the the Hamiltonian contributes 1

2kT

to the mean energy.1

1 2

L

RR

Z=R

FIG. 4. Lossless transmission line Z = R of length L withmatched terminations.

The energy density per unit frequency U(ω) is thengiven by the product of the density of modes and themean energy per mode:2

U(ω) = D(ω)〈ε(ω)〉

=kT

2πv(4)

The power per unit frequency is then simply:3

P (ω) = vU(ω)

=kT

2π(5)

OR

P (ν) = kT (6)

This is the power per unit frequency absorbed by the re-sistor. By the principle of detailed balance this must beequal to the power per unit frequency emitted by the re-sistor. The thermal electromotive force generated by theresistor sets up a current I = V/2R in the transmissionline. Thus, the power absorbed by the resistor at theother end is

P (ν) = 〈I2(ν)〉R (7a)

=

⟨V 2(ν)

4R2

R (7b)

=〈V 2(ν)〉

4R(7c)

Equating Eq. 6 to Eq. 7c and then solving for themean-square voltage per unit frequency gives:

〈V 2(ν)〉 = 4RkT (8)

By integrating the expression above over the accesiblefrequency range, we arrive at the Nyquist Theorem:

〈V 2〉 = 4kTR∆ν (9)

1The Hamiltonian (per unit volume) for an electromagneticwave is given by H = 1

8π(E2 +B2).

2U(ω) is a one-dimensional energy density.3Recall that the energy density is equivalent to a force.

2

Page 3: Johnson Noise - Semantic Scholar · mal noise, the Boltzmann constant was measured to an accuracy of»4%.7 Alternatemethods implementing a undergraduate physics experiment on Johnson

B. Microscopic Derivation

Consider a conductor of resistance R with a charge car-rier densityN having a relaxation time τc. The conductorhas length ` and cross-sectional area A. The voltage Vacross the conductor is

V = IR (10a)

= RAj (10b)

= RANe〈u〉 (10c)

where I is the current, j is the current density, e is thecharge on an electron, and 〈u〉 is the drift speed alongthe conductor.

Noting that NA` is the total number of electrons inthe conductor,

i

ui = NA`〈u〉 (11)

Solving for 〈u〉 in Eq. 11 and substituting the resultingexpression into Eq. 10c gives,

V =∑

i

Vi =Re

`

i

ui (12)

where ui and Vi are random variables.The spectral density J(ν) has the property that in the

frequency interval ∆ν

〈V 2i 〉 = J(ν)∆ν (13)

The correlation function can be written as

C(τ) = 〈Vi(t)Vi(t+ τ)〉 (14a)

= 〈V 2i (t)〉 exp (−τ/τc) (14b)

where τ is an arbitrary time interval.By substiting Eq. 14b and Eq. 12 into the Wiener-

Khintchine theorem Eq. 15a, the spectral density is

J(ν) = 4

∞∫

0

C(τ) cos (2πντ) dτ (15a)

= 4

(Re

`

)2

〈u2〉

∞∫

0

exp(−τ/τc) cos (2πντ) dτ (15b)

= 4

(Re

`

)2

〈u2〉τc

1 + (2πντc)2(15c)

≈ 4

(Re

`

)2

〈u2〉τc (15d)

≈ 4

(Re

`

)2(kT

m

)

τc (15e)

where 〈u2〉 = kT/m by the equipartition theorem. Notethat for metals at room temperature τc < 10−13, thusfrom the DC through the microwave range 2πντc ¿ 1.

Thus the mean-square voltage in the frequency range∆ν equals:

〈V 2〉 = NA`〈V 2i 〉 (16a)

= NA`J(ν)∆ν using Eq. 13 (16b)

= NA`4

(Re

`

)2(kT

m

)

τc∆ν using Eq. 15e (16c)

= 4

(Ne2τc

m

)A

`R2kT∆ν (16d)

Using a result from conductivity theory σ = Ne2τc/mand the elementary relation R = `

σA [5]:

〈V 2〉 = 4 σA

`︸︷︷︸

1/R

R2kT∆ν (17)

We have once again arrived at the Nyquist Theorem:

〈V 2〉 = 4kTR∆ν (18)

The Nyquist Theorem is a special case of the general con-nection existing between fluctuations (random variables)and dissipation in physical systems. Brownian motionlends itself to a similar analysis [6], [7].

III. EXPERIMENTS

This section describes the experimental apparatusused, the calibration performed and the measurementsthat were recorded.

A. Apparatus

Figure 5 is a diagram of the experimental apparatusused to measure the Johnson Noise.4 An inverted beakershielded the resistor R which was mounted on the ter-minal of the aluminum box. The resistor is connectedto the measurement chain through two switches (SW1and SW2). A Hewlett-Packard HP54601A digital oscillo-scope was used to measure the root-mean-square (RMS)voltage generated by the resistor. Because the JohnsonNoise signals are in the microvolt range, a low-noise am-plifier (PAR 113) was used to produce millivolt signalsdetectable by the digital oscilloscope. A band-pass filter(Krohn-Hite 3202R) was used to prevent thermal noiseoutside the frequency range 1 KHz – 50 KHz from be-ing amplified.5A Tektronix Function Generator (FG) 504

4Figure 5 was scanned-in from the junior lab guide [8].5Signals outside this frequency range could not be properlyamplified by the PAR 113.

3

Page 4: Johnson Noise - Semantic Scholar · mal noise, the Boltzmann constant was measured to an accuracy of»4%.7 Alternatemethods implementing a undergraduate physics experiment on Johnson

provided sinusoidal calibration signals. The FG and theKay attenuator were used to calibrate the measurementchain.

Several steps were taken to filter out extraneous noisefrom the experimental apparatus. At all times the digi-tal oscilloscope was kept at least five feet from the noisesource, otherwise the variable magnetic field from itsbeam-control coil would produce undesirable electricaloscillations in our noise measurements. Coaxial cableswere also kept as short as possible to keep minimize elec-trical interference.

FIG. 5. Experimental apparatus.

B. Calibration of Measurement Chain

1. Test signal RMS voltage

The amplitude of the sinusoidal signal produced fromthe FG was adjusted so that the RMS voltage VRMS asmeasured on the digital oscilloscope was approximately 2volts. The RMS voltage of the FG sinusoidal signal wasrecorded over the range passed by the Krohn-Hite Filter(refer to Figure 6). It was confirmed that the RMS volt-age varied slightly over the frequency range of interest.

2. Gain of measurement chain

The sinusoidal test signal was fed through the Kay at-tenuator set to 60 dB (1000) of attenuation to the ‘A’input of the PAR amplifier (set to 1K) with ‘B’ inputgrounded. The RMS voltages out of the Krohn-Hite fil-ter were measured over a 100 kHz frequency range. Thegain squared [g(ν)]2 was small at very low frequency, thendrastically increased to unity around 5 kHz (refer to Fig-ure 7). As expected at higher frequency (> 50 kHz) thegain squared roll off considerably.

0 10 20 30 40 50 60 70 80 90 1001.98

1.985

1.99

1.995

2

2.005

2.01

ν (KHz)

VR

MS (

V)

Variation of test signal VRMS

with frequency

FIG. 6. RMS voltage VRMS produced by function generatoras a function of frequency.

C. Resistance Dependence of Johnson Noise

With the PAR amplifier set to 1K, typical RMS volt-ages out of the Krohn-Hite filter were in the millivoltrange. The component of the noise VS not generated bythe resistor but by the amplifier itself was measured by:

1. Opening SW2.

2. Unplugging the connections to the ohmmeter andtemperature meter.

3. Shorting the resistor with SW1.

The total RMS voltage VR was measured with the short-ing switch SW1 open. Because all the contributions tothe measure RMS voltage are statistically uncorrelated,they add in quadrature. Thus, mean square Johnson noiseof the resistor is given by,

V ′2Jo = V 2

R − V 2S (19)

where VR and VS are the RMS voltages measured withthe SW1 open and closed, respectively. The resistance Rwas measured using a digital multimeter after each noisemeasurement.

D. Temperature Dependence of Johnson Noise

The Johnson noise of a 22.2 kΩ resistor was measuredat liquid N2 temperature −160C to 150C. High temper-atures were obtained by mounting the inverted aluminumbox and placing it on a cylindrical oven. The tempera-ture was adjusted by using a Variac. Low temperatures

4

Page 5: Johnson Noise - Semantic Scholar · mal noise, the Boltzmann constant was measured to an accuracy of»4%.7 Alternatemethods implementing a undergraduate physics experiment on Johnson

−20 0 20 40 60 80 100 1200

0.2

0.4

0.6

0.8

1

1.2

1.4

ν (KHz)

[g(ν

)]2

Gain squared of the measurement chain as a function of frequency

FIG. 7. Gain squared of measurement chain in the fre-quency range (0.5 KHz – 100 kHz.) NOTE: The dotted lineis not a fitted function. Its purpose is tom emphasize a trendin the gain squared. The gain squared has been normalizedsuch that the value of [g(ν)]2 = 1 corresponds to a gain of1000.

were obtained by inverting the aluminum box and plac-ing it on a liquid N2 filled dewar flask. The temperaturewas varied in an ad-hoc manner by raising and loweringthe aluminum box into the dewar flask as needed.

IV. RESULTS AND DISCUSSION

The first subsection explicitly connects the NyquistTheorem with the experimental setup at hand. The lasttwo subsections describe the results of the subsectionsIII C & IIID, respectively.6

A. Derivation of RMS thermal voltage at the

terminal of an RC circuit

The resistor and coaxial cables that are connected tothe PAR amplifier can be modeled as the circuit depictedin Figure 8. The equivalent circuit is composed of a fluc-tuating thermal electromotive force VJo with an ideal re-

6Note that Boltzmann constant is calculated in the last two

subsections.

sistor R and a capacitor C in a simple lowpass filter con-figuration.

C

R

V’Jo

VJo

FIG. 8. Equivalent circuit of the electromotive force acrossa conductor of resistance R connected to the measuring devicewith cables having capacitance C.

In sinusoidal steady state, impedances can be used totreat the circuit as a voltage divider.

V ′

Jo =(iωC)−1

(iωC)−1 +Rg(ω)VJo (20a)

=1

1 + iωCg(ω)VJo (20b)

The RMS thermal voltage is the magnitude of Eq. 20b:

V ′2Jo =

[g(ν)]2V 2Jo

1 + (2πνRC)2(21)

The Johnson Noise is equation Eq. 21 summed over theaccessible frequencies,

V ′2Jo = V 2

Jo

∞∫

0

[g(ν)]2

1 + (2πνRC)2

︸ ︷︷ ︸

G

dν (22)

In this experiment, the integral in Eq. 22 was numericalevaluated using the data collected in the calibration ofthe measurement chain (Figure 7). The capacitance Cwas approximated at 60 pF from considerations of theamount of coaxial cable used and its known capacitanceper unit length, 30.8 pF/feet. The Nyquist Theorem ex-pressed in terms of the present variables is arrived atby taking Eq. 9(or 18) and making the substitutions:〈V 2〉 → V ′2

Jo and ∆ν → G.

V ′2Jo = 4kTRG (23)

B. Determination of the Boltzmann Constant

The RMS voltage was measured for eight metal filmresistors (whose values ranged from 20 kΩ to 103 kΩ) atroom temperature. Figure 9 is a plot of V ′2

Jo against R.The Boltzmann constant was calculated by solving for kin Eq. 23. The experimentally determined value of theBoltzmann constant, 1.37± 0.06× 10−23 J/K, is in goodagreement with the accepted value 1.38× 10−23 J/K.

5

Page 6: Johnson Noise - Semantic Scholar · mal noise, the Boltzmann constant was measured to an accuracy of»4%.7 Alternatemethods implementing a undergraduate physics experiment on Johnson

0 200 400 600 800 1000 1200−1

−0.5

0

0.5

1

1.5

2x 10

−12

R (kΩ)

V2 Jo

’/R (

V/k

Ω)

RMS voltage for Johnson Noise for a variety of resistors

FIG. 9. Resistance dependence of Johnson Noise V ′Jo.

C. Determination of the Absolute Zero on

Centigrade Scale

The RMS voltage for 22.2 kΩ resistor was measuredat fourteen temperatures ranging from ∼ −160C to∼ 150C at approximate intervals of 25C Figure 10 is aleast-squares fit of V ′2

Jo/4RG vs. T . The slope of the linegives the Boltzmann constant and the T -intercept is thecentigrade temperature of absolute zero. The Boltzmannconstant was determined to be 1.363±0.025×10−23 J/Kand centigrade temperature of absolute zero was extrapo-lated to −265.5±6.9C. Both experimentally determinedvalues are in good agreement with their accepted valuesof 1.38× 10−23 J/K and −273.15K, respectively.

−200 −150 −100 −50 0 50 100 150 2000

1

2

3

4

5

6

7x 10

−21 Temperature dependence of Johnson Noise

V2 Jo

’/4G

R (

V k

Ω−

1 s

ec−

1)

T (°C)

FIG. 10. Temperature dependence of Johnson Noise V ′Jo.

V. CONCLUSIONS

Johnson Noise belongs to a broader category ofstochastic phenomena which have been of research in-terest for decades. Measurement of the thermal noisein resistors provided a means to calculate the Boltzmannconstant and the centigrade temperature of absolute zero.Because there are inherent difficulties in measuring ther-mal noise, the Boltzmann constant was measured to anaccuracy of ∼ 4 %.7 Alternate methods of implementinga undergraduate physics experiment on Johnson Noiseare described in the literature (e.g. [9]).

ACKNOWLEDGMENTS

I would like to thank Mukund T. Vengalatorre for hisassistance in carrying out this experiment. I acknowl-edge Dr. Jordan Kirsch for the many useful discussionson the subject of thermal noise. The author is gratefulto Cyrus P. Master for editing a preliminary version ofthis document.

[1] H. Nyquist, Phys. Rev. 32, 110 (1928).[2] F. Reif, Fundamentals of Statistical and Thermal Physics

(McGraw-Hill, New York, 1965), pp. 567-600.[3] J. B. Johnson, Phys. Rev. 32, 97 (1928).[4] C. Kittel, Elementary Statistical Physics (Wiley, New

York, 1967), pp. 147-149.[5] E. M. Purcell, Electricity and Magnetism (McGraw-Hill,

New York, 1985), pp. 138.[6] G. E. Uhlenbeck and L. S. Ornstein, Phys. Rev. 36, 823

(1930).[7] S. Chandrasekhar, Revs. Mod. Phys. 15, 1 (1943).[8] G. Clark and J. Kirsch, Junior Physics Laboratory: John-

son Noise, M.I.T., Cambridge, Mass., 1997.[9] P. Kittel, W. R. Hackelman, and R. J. Donnelly, Am. J.

Phys. 46, 94 (1978).

7In his original paper, Dr. J. B. Johnson measured the Boltz-mann constant within 8 % of the accepted value.

6