Sunteți pe pagina 1din 29

Establishing a Hall Potential in Silver

via an Orthogonal Magnetic Field to


Deduce the Density of Accumulated
Negatively Charged Particles

Created by Brian Hallee

Partnered by Joseph Oxenham

Performed October 29, 2010


Historical Background
The Hall Effect is known to be one of the most precise and defining

physical phenomena to ever be discovered. Its discoverer, of whom the

effect has since been named after, was Edwin Hall who happened across it

while performing experimental work for his doctoral dissertation in 1879.1

Over the course of his doctoral studies, which naturally took place in the

realm of physics, Hall experimented with small metal leaflets by passing a

current through their length and a magnetic field through their face

(orthogonally). After subsequently tapping off the leafs at various points

down their length with a voltmeter, he noticed that in a short amount of time

that the presence of a magnetic field would cause a very small potential

difference from top-to-bottom. This harmoniously further united electric

current with magnetic fields, and lead 19th century physicists further down

the road to understanding the quantum mechanical atom.2 An interesting

setback to the results of this experiment was the fact that the electron was

still a mystery at this point in time. J.J. Thompson of Cambridge University

would not go on to rigorously prove its existence another two decades.

Thus, Hall and others had to await the formation of quantum mechanics in

order to fully conceptualize his discovery. While the significance of this

seemingly trivial electric potential will be developed and explained fully in

Page | 2
the theory section, it is worth noting that Hall’s discovery has spawned

several related discoveries on the quantum scale.3 These new quantum-

based Hall effects have not only found their way into many modern

applications you likely use every day, they have also helped to narrow down

physical constants (i.e. the fine structure constant) to one part in a billion.4

Only three decades ago was the Hall Effect generalized to the quantum

scale. The “integer” quantum hall effect was first predicted by three

physicists in 1975 who, in turn, doubted their own calculations. Nonetheless,

five years later, a team led by Dr. Klaus von Klitzing discovered that the Hall

conductivity increased by, what physicists now grant as, exact integers. 5

Considering this is rather beyond the scope of this report, I ask you to treat

as an axiom the fact that this effect has helped narrow the value of he2 and

the value of the “fine structure constant” α. (A value of utmost importance

in quantum electrodynamics) The fraction he2 has since been bestowed the

term “von Klitzing Constant” in his honor, and in 1985 the quantum Hall

effect had become relevant enough to grant him the Nobel Prize in Physics of

that year.6 With the advancement of integrated circuit technology, the Hall

Effect has found useful application in the form of Hall Probes or Hall Sensors.

These devices often serve as miniature magnetometers mostly for magnetic

field detection and flux leakage measuring. Hall Probes require integrated

circuit amplifiers for the reason that, as we will observe in the following

sections, Hall voltage levels are extraordinarily small and require large

Page | 3
sophisticated voltmeters for accurate measurements at this scale. Typically,

they find a home in all types of sensing equipment such as speed, pressure,

current, or fluid-flow sensing. They are also likely to be paired up with a

potentiometer for applications that require a switch to be robust (eg. electric

guns).7 While one would be inclined to believe that physicists would have,

more or less, “closed the book” on the Hall Effect by stretching and

generalizing it as far as is physically feasible over the past 100 years, this is

not the case. Currently, heavy research is involved in understanding the

physics behind the fractional quantum Hall Effect; discovered in 1982. As

common sense would suggest, the discovery of this phenomenon stemmed

from the observations of Hall steps with fractional quantum numbers. The

brunt of the understanding of this occurrence comes from the utilization of

“wave-functions”. However, as this article undergoes construction, heavy

research remains in the realm of “quasi-particles” and their atypical

fractional charge.

Theoretical Basis

Fortunately, as scientists of

the twenty-first century, we

are able to apply the electron

and proton to our conceptual

examples to fully grasp the

Page | 4

Figure 1: A semiconductor experiencing the Hall


Effect
phenomenon and apply mathematics to it. The derivation of the hall voltage

is relatively simple an only requires a firm understanding of current,

magnetic force, and Newton’s second law. We will refer to figure 1 quite

frequently, and it is only fitting that we inform you of the meaning behind the

variables it contains at the start.

I = The current passing through the semiconductor (Natural convention calls


for this to designate positive charge movement
(protons) {Depicted in the diagram} We will observe shortly that we are only
concerned with electron movement in this experiment.
Fm = The magnetic force exerted on the moving charge carriers
B = The magnetic field
dT = The thickness of the plate (In our derivation, we will simply designate
this as T)
D = The height of the plate
FE = The electric force
VH = The Hall Voltage. (To differentiate this from regular potential and
velocity, we will use Hv)

If we further bestow another variable to the figure, dX (An infinitesimally

small length spanning the “height” of the block), we can state that a charge

dQ moves through the length dX over a time dt. Unfortunately for Hall, this

is where his toils would have ended, as we are ready to apply the notion of a

“charged particle”, or electron, to the derivation. At this point, the volume

dX contains an amount of charge dQ, or the number of individual charged

particles times the volume they occupy. Thus,

Charge in dX =dQ = ndV (eqn. 1)

We can infer from figure one that the volume dV is equal to T*D*dX.

Therefore,

Page | 5
dQ=nTD*dX (eqn. 2)

If we divide equation 2 by the differential time we spoke of earlier, we

obtain:

dQdt = nTDdXdt → I = nTDvd (eqn. 3)

We utilized the fact that charge per-unit time is current, and distance per-

unit time is velocity to achieve equation 3. The subscript d for velocity

denotes the “drift” nature of this movement. Barring unnecessary

electrodynamics, the drift velocity arises from the mostly random motion of

electrons. Although we tend to believe that all electrical phenomena operate

at relativistic speeds, this is simply not the case for mean electron

movement in a wire. When an electron field is present, they will tend to drift,

on the order of 10-4m/s, towards their destination due to the large amount

“roadblocks” (nuclei, inter-molecular forces, etc.) they must overcome. 8 This

drift velocity will come in handy as our derivation progresses. Thus we re-

write equation 3 as follows:

vd = InTD (eqn. 4)

Next, we consider the magnetic field and its effects on our stream of charged

particles. The general equation governing magnetic force on a charged

system is written as follows:

FB = q*v x B (eqn. 5)

Page | 6
Before we move any further, we must denote what charges are actually of

focus in our experiment. To solve this, we will treat as an axiom the fact that

semi-conductors (such as Silver, our metal of choice) allow for negative

charge movement within their structure.9 In layman’s terms, when a current

develops in our sample, we can safely assume that electrons are the cause

of this, and they move in the opposite direction of standard convention.

Thus, when applying equation 5 both quantitatively and qualitatively, we

must take into account the negative charge attributed to electrons. We will

begin with the qualitative argument to support figure 1 and its, thus far,

mysterious charge accumulations.

FB = - -i x -k = +j (eqn. 6)

Equation 6 is strictly concerned with the vector analysis of equation 5. It

predicts that an electron with velocity moving in the opposite direction of the

current (denoted (-i)) acted upon by a magnetic field perpendicular to its

velocity (denoted (-k)) will be forced “upward” along the height of the metal

plate. By this reasoning, it seems as though in a short time negative

electrical charge will accumulate towards the top, and this is exactly what we

observe experimentally. In fact, this phenomenon naturally causes a net

positive charge accumulation on the bottom and, thus, a Hall Potential is

born between the two. This is precisely what Dr. Hall observed, and we will

Page | 7
continue our quantitative derivation to arrive at a clean formula to predict

what this potential might be. Taking the absolute value of equation 5:

|FB| = qvB = -evdB (eqn. 7)

Thus, we observe the drift velocity coming back into play, and we require

one more observation to tie it all together. Once the accumulations have

settled down, a miniscule net electric field develops between the positive

end and the negative end. From introductory electrostatics, we know:

FE = q*E = -e*E (eqn. 8)

Allotting enough time for the accumulations to reach equilibrium, (In reality,

this takes only seconds), we are able to apply Newton’s first law to the

electric and magnetic forces:

FE = -FB (eqn. 9)

Therefore,

eE = evdB (eqn. 10)

Electric field has units volts/meter. Thus, again referring to figure 1, we can

set E= HvD. Equation 10 becomes:

HVD = vdB (eqn. 11)

Page | 8
Finally, substituting in equation 4 and multiplying by D, we arrive at the Hall

Voltage formula:

Hv = DIBnTD = InT*B (eqn. 12)

While equation 12 looks about as clean as clean gets, therein lies a problem

with the term in parenthesis. N can, by no means available today, be

measured directly or even guessed. Typically, there are somewhere on the

order of 1010 electrons that accumulate on the surface of a semiconductor at

low currents. However, while we mentioned earlier that direct measurement

of Hall Potentials require rather precise, high-quality equipment, we were

able to utilize such gear in our attempt at this lab. Naturally, magnetic field

is easily measureable with a standard gauss meter. Thus, our notation used

in equation 12 has become clear, as InT represents the slope of two

measurable quantities. The most efficient way to solve for n, our ultimate

goal, is to measure the Hall Potential at different current and field strengths

and obtain the slope of the subsequent graph. From this, the charge carrier

density in the accumulation can be found using simple algebra:

n = Islope* T (eqn. 13)

To place n into perspective, electrons wield a mass on the order of 10-31kg

while the mass of silver roughly 0.5 kg. Thus, we will expect the density of

particles in the accumulation to be very large.

Page | 9
Apparatus

While the

theory behind

the Hall

Potential may

have come

across as

relatively

simple, the

equipment used

to
Figure 2: The Walker Scientific Electromagnet

experimentally achieve it was not. This lab utilized a slew of meters, a

standalone, dedicated, water-cooled, DC power source, and a 3-Tesla Walker

Scientific electromagnet as seen in figure 2. Starting with the meters, we

were required to monitor the Hall Voltage, current, and field strength

simultaneously. Thus, we were given a voltmeter that read voltage to the

order to 10-5 volts and auto-compensated jitters, a battery-operated

Page | 10
ammeter, and a gaussmeter. All were pre-assembled and pre-calibrated as

we arrived to carry out the lab procedure. A photo of the meters can be

viewed in figure 3 below. The power supply played an important role in our

ability to follow through with this lab. As noted in the lab handout9, at

maximum current the electromagnet consumes 8500 Watts of power. At this

level of consumption, we were required to cool and operate a DC power

supply coupled to an entirely different breaker from the rest of the room. The

cooling water stemmed from the local watershed, and was used to cool both

the DC power supply and the electromagnet. Also pictured in figure 3 is the

current generator used to move charge carriers through the silver plate.

Lastly, the device of interest (Silver plate) can be viewed in figure 2 centered

between the North and South poles of the electromagnet. The North end of

the magnet is located on the left, while the South resides on the right. Thus,

the orthogonal magnetic field points from left to right (North-to-South) in

figure 2. Meanwhile, the conventional current flow moves from back to front

Page | 11
Figure 3: (From left to right) Voltmeter, Ammeter, Current source, Gaussmeter

in the Hall device. Consequently, we should expect the negative charge

accumulation to reside at the top of the device. We shall return to this topic

in the discussion section when we hypothesize positive charge carrier

movement and its effects.

Procedure

Due to the very sensitive, (and not to mention expensive), nature of

the equipment present for this experiment, the procedure we followed was

rather rigorous and static. In other words, the only acceptable methodology

to use in performing this experiment was to follow that stated in the lab

report. In order to ensure proper startup of the electromagnet, Dr. Greg

Latta (Instructor, Professor of Physics) made a brief appearance to oversee

our procedure. To begin, we switched on the power strip that served the

voltmeter, ammeter, and gauss meter and followed this by switching on the

three respective meters. Next, we switched on the DC power supply and the

water coolant that served it. Before doing this, however, it was of utmost

importance that the coarse, medium, and fine knobs were all set to zero on

the power supply. This is due to the fact that a sudden increase or decrease

in power to the electromagnet could cause severe damage to the internal

components. As we activated the water coolant, the power supply required

monitoring until a light designating sufficient water flow was galvanized.

Once this occurred, we were able to switch on the current supply and set it

Page | 12
to 5.0 Amps. At this point the voltmeter displayed a relatively large potential

even though no current was passing through the Hall device. We will return

to this in the discussion section, but for now it is sufficient to say that

adjusting the Hall device and auto-compensating the “jitter” on the

voltmeter was adequate to begin our data gathering. The actual data

acquiring section of the lab simply required the magnetic field to be

increased in increments of 0.1 Tesla. At each increment we entered the Hall

Potential into our pre-constructed Microsoft Excel spreadsheet. This

spreadsheet enabled us to view the linear curve between Hv and B in real

time. We divvied the tasks amongst the two of us. While I was responsible

for reading the Hall Potential and entering the data into Excel, Joseph

operated the DC power supply. Even after compensation, jitter was still

present in the Hall Potential. Thus, we were required to average the value

on the fly, as waiting too long to gather a reading could potentially introduce

negative effects. Also worth noting, the Voltmeter was set to a scale of 10-
5
V. Hence, we were forced to quickly convert the measurement into

microvolts by multiplying the raw data by ten. We continued this trend until

reaching a magnetic field of two tesla. At this point, we brought the DC

power, and the subsequent field, back down to zero. We repeated this

process again for a current of 10 Amps, and again at 15 Amps. A new auto-

compensation was required between each run. At the conclusion of the lab,

Page | 13
we ensured that both the current and DC power was set to zero. At this

moment, all devices were turned off and the water coolant was stoppered.

Sample Calculations

Figure 4: The graph resulting from our first run with a current of 5 Amps
(Refer to the accompanying spreadsheet for the graphs pertaining to runs 2 & 3)

Figure 4 depicts the graph we achieved from the data of our first run.

Looking at our coefficient of determination, we see that our data almost

perfectly fits the linear relationship as predicted by equation 12. I feel as

though it is important, at this point, to express the expected value for the

charge carrier density in units of electrons-per-meter cubed in order to

compare sample values to it. That value is: 6.9x1028 em3. You may view our

experimental values for charge carrier density in the appendix of this report.

The formula we utilized to achieve those values is shown below:

ne=B6SLOPEB8:B27,A8:A27*$B$4*1061.6 x 10-19 (1)

We wish to check this algorithm by performing a sample calculation of our

own. In this example I will use the data points B = 1.0T and B = 1.1T from

our first run using I = 5A. Thus,

9.2μV=m*1.0T+B & 10.0μV=m*1.1T+B

Where m denotes the slope of the line formed by this data and B represents

the y-axis intercept. If we subtract the two equations, we achieve:

Page | 14
0.8μV=m*0.1T → m= 0.8μV0.1T = 8μAm2

Looking at figure 1, we see that our slope was 8.2 μAm2, so we can be sure

our methodology is correct so far. Next, we use the data point B = 1.0T and

our newfound slop to solve for B:

9.2μV=8μAm2*1.0T+B → B = 9.2μV8μA*Tm2 = 1.15μV

Again, this closely matches our actual value for the intercept of 1.1311μV.

While B was found primarily for justification, the slope is where our interests

lie. We can now apply equation twelve to the slope and attempt to attain a

value for the carrier density using two data points:

slope= 8μAm2 = InT (2)

We already know the current as it was set to roughly 5 amps. Likewise, T

(plate thickness) is a constant given in the lab handout to be 5.41x10-5m.

Thus, we rearrange (2) using equation 13 to solve for n as follows:

n = Islope * T = 5A8μAm2* 5.41*10-5m = 1.55x104μCm3 (3)

In this state, the charge carrier density doesn’t really tell us much. Our goal

is to find the amount of electrons per meter cubed. First, we multiply (3) by

106 to achieve a value in coulombs:

n = 1.55x104μCm3 *1C10-6μC = 1.55x1010Cm3 (4)

Page | 15
Lastly, we apply the fact that one electron exhibits a charge of 1.60x10-19C.

Thus, we divide (4) by this value to achieve our sought after value in units of

em3:

ne = 1.55x1010Cm3*e1.60x10-19C = 7.22x1028 em3

Considering the massive number of electrons accumulating on this piece of

silver, it seems rather amazing that we come so close to the expected value

with only two data points. We can perform a sample error calculation

relative to the expected value shown below:

% error = | n0 - ne |n0*100% = 6.9 - 7.226.9*100% = 4.64%

This is an almost insignificant error considering the sizes of the values. Thus,

as we have demonstrated over the previous two pages, we can be very sure

that our excel spreadsheet algorithms, graphs, and final values are correct

and within reason.

Discussion

Upon averaging the values for ne acquired after three runs, we

achieved a final value of:

Page | 16
ne = 7.07x1028 em3

As stated in the previous section, the accepted value of ne, according to the

lab handout, was 6.9x1028 em3. Consequently, we calculate our relative

error as we did before (The factors of 1028 all cancel out):

% Relative Error = 6.9 - 7.076.9*100% = 2.46% error

This is a rather exceptional error, and, after reviewing our stellar coefficients

of determination and our lack of any setbacks whatsoever during the course

of our lab, truly wraps up a successful experiment. This error comes off as

insignificant due to the fact that the size of our “sample” is on an order to

which nothing else on earth compares to (other than more atoms). Even the

farthest galaxy currently known, Abell 1835 IR191, is only 13 billion light

years (1.3x1010) away.10 Our density is more than a billion billion times

larger that that distance. Thus, the fact that in half an hour we were able to

achieve a roughly 2% error on this quantity simply with a volt and gauss-

meter is simply extraordinary.

As we mentioned in the procedure section, we were forced to auto-

compensate for large potentials that would accumulate before any current

was activated. This phenomenon is known as thermo-couple potential

creation, and it is due to the temperature gradient between the junctions

located throughout the device. When there is temperature difference

Page | 17
between two conductors two important phenomena occur: Heat flow, and

what is coined as the Seebeck effect.11 We are only concerned with the

latter. The Seebeck effect concerns the fact that energetic electrons will

move from the hot junction to the cooler one while pushing some of the

lesser energy electrons with them. In turn, this causes a potential difference

between the hot and cold ends, and, alas, a thermocouple potential is born.

Considering the fact that not only did we perform this lab early in the

morning, but we were the first to perform it that day, we likely had to

compensate for more of this potential than one normally would. Likewise,

this may have been a source of some of the error we did garner as the

junctions and conductors heated up over the course of the experiment. You

might notice that our first run returned a value a bit lower than the other

two. I postulate that this is due to the heating of the wires over the first run,

in turn, throwing off our initial

compensation.

Next we consider the situation in

which positive charge carriers would

have moved through the Hall Device.

Technically, this effect transpires

from the movement of “holes” in the

metal; not protons. However, for

simplicity, we will use the term


Page | 18

Figure 5: The Hall Effect Device


“protons” in place of positive charge carriers. If we assume the protons to

be moving from the back of the Hall Device (see figure 5) to the front, then,

using the same directional convention as before, the velocity of the protons

has a directional vector in the positive x, or i, direction. The magnetic field

does not change. Thus, it continues to move from North to South in the

negative z, or -k, direction. We will again make use of equation 6 to

qualitatively observe the physics of positive charge movement. Summing up:

vprotons = i & B = -k & q= +e

Therefore:

FB =+*i x -k= +j

By this prediction, if protons truly are the charge carriers moving through the

Hall Device, the polarity of the charge accumulation is reversed, as positive

charge now occupies the top of the silver slab. While, theoretically, the

absolute value to the Hall Potential should remain the same, the field is

reversed as protons move through the metal.

Further proof that thermocouple potentials likely played a role in our first run

is the peculiar jump in our value for ne between the first run and the other

two. Runs two and three returned almost the exact same value, while the

first run, although closer to the expected value n0, was the odd ball. At first

glance, we figured that this was likely due to an increased current, which

Page | 19
subsequently allowed for a greater amount of electrons to be pushed to the

top. However, looking more carefully at equations 12 and 13, while the slope

varies linearly with current, the value of ne has an inverse relationship with

this increased current. Thus, in theory, the increase in current should be

cancelled out by the increase in slope, and ne should remain constant.

Therefore, returning to our original prediction, we feel as though the

variance in our first run was due to a cold apparatus exhibiting thermocouple

potentials.

One interesting phenomena we were unable to experiment with in our lab

session is the moving of the conductor itself. Naturally, equation 6 is only

concerned with the movement of the electrons relative to the electro magnet

(or, specifically, the electric field). Thus, we can hypothesize that if we were

to move the entire Hall Device in the opposite direction of the current (in our

case, toward the back of the electromagnet) at the exact speed as the

electron drift speed, then the motion of the electrons would be zero relative

to the field. If this is the case, then equation 6 zeros out and we have a

magnetic force of exactly zero. Succinctly, if we move the Hall device at an

equal and opposite velocity relative to the electrons inside, we will not

observe a Hall potential! Achieving this even near-perfectly in the lab setting

would be rather difficult. However, the theory further solidifies the relation

between magnetic fields and charge carriers.

Page | 20
Although 19th century scientific advancement hinders us from directly

comparing our value of electron density with Hall’s, we can qualitatively

appreciate the rigor involved in his finding of such a remarkably small

potential difference. Nonetheless, the effect has found applications far and

wide due to its ability to give a clean on-off signal. While the classical Hall

Effect, at this point, may be fully understood, its extensions into the quantum

world will likely be a topic of heated debate and research for years to come.

Page | 21
Works Cited
Cain, F. (2004, March 1). Record for Furthest Galaxy is Broken Again.
Retrieved October 30, 2010, from Universe Today:
http://www.universetoday.com/9347/record-for-furthest-galaxy-is-
broken-again/

Georgia State University. (n.d.). The Hall Effect. Retrieved October 30, 2010,
from HyperPhysics: http://hyperphysics.phy-
astr.gsu.edu/hbase/magnetic/hall.html

Honeywell. (n.d.). Hall Effect Sensing and Application. Retrieved October 30,
2010, from Honeywell:
http://content.honeywell.com/sensing/prodinfo/solidstate/technical/hall
book.pdf

Hugh D. Young, R. A. (2007). University Physics. Pearson Addison-Wesley.

Latta, D. G. (n.d.). The Hall Effect in Silver Notes. Frostburg: Frostburg State
University.

Microstar Laboratories. (2009). Thermocouple Cold Junctions. Retrieved


October 30, 2010, from Microstar Laboratories:
http://www.mstarlabs.com/sensors/thermocouple-cold-junctions.html

Qiu, Y. (1997, April 27). Basics of Hall Effect: History. Retrieved October 30,
2010, from Johns Hopkins University:
http://www.pha.jhu.edu/~qiuym/qhe/node1.html

Page | 22
ENDNOTES

Page | 23
APPENDIX - RAW DATA

Current: 4.98 A Current: 10.01 A Current: 15.01 A

B (T) Hv (μV) Reading B (T) Hv (μV) Reading B (T) Hv (μV) Reading

0.1 2.00 0.20 0.1 1.60 0.16 0.1 2.50 0.25

0.2 2.90 0.29 0.2 3.00 0.30 0.2 5.30 0.53

0.3 3.70 0.37 0.3 4.60 0.46 0.3 7.40 0.74

0.4 4.50 0.45 0.4 6.20 0.62 0.4 9.60 0.96

0.5 5.30 0.53 0.5 7.80 0.78 0.5 12.50 1.25

0.6 6.00 0.60 0.6 9.60 0.96 0.6 14.90 1.49

0.7 7.00 0.70 0.7 11.10 1.11 0.7 17.50 1.75

0.8 7.70 0.77 0.8 12.90 1.29 0.8 19.80 1.98

0.9 8.30 0.83 0.9 14.10 1.41 0.9 22.30 2.23

1.0 9.20 0.92 1.0 16.10 1.61 1.0 24.50 2.45

Page | 24
1.1 10.00 1.00 1.1 17.60 1.76 1.1 27.10 2.71

1.2 10.80 1.08 1.2 19.20 1.92 1.2 29.60 2.96

1.3 11.70 1.17 1.3 21.00 2.10 1.3 32.00 3.20

1.4 12.70 1.27 1.4 22.50 2.25 1.4 34.20 3.42

1.5 13.50 1.35 1.5 24.40 2.44 1.5 36.80 3.68

1.6 14.40 1.44 1.6 25.90 2.59 1.6 39.40 3.94

1.7 15.30 1.53 1.7 27.40 2.74 1.7 41.90 4.19

1.8 16.10 1.61 1.8 29.20 2.92 1.8 44.10 4.41

1.9 16.80 1.68 1.9 30.50 3.05 1.9 46.50 4.65

2.0 17.60 1.76 2.0 32.10 3.21 2.0 48.60 4.86

1.1182E+ 11387638 1.14E+


n= 10 C/m^3 n= 878 C/m^3 n= 10 C/m^3

6.9886E+ 7.11727E 7.11E+


ne = 28 e/m^3 ne = +28 e/m^3 ne = 28 e/m^3

ne-avg 7.07315E
= +28 e/m^3

Page | 25
Page | 26
APPENDIX - DISC CONTENTS
• ROOT DIRECTORY

○ LAB 4 – HALL EFFECT.DOC


 THE OFFICIAL MICROSOFT WORD LAB REPORT CONCERNING THE HALL EFFECT IN

SILVER

○ HALL EFFECT SPREADSHEET.XLS


 THE MICROSOFT EXCEL SPREADSHEET CONTAINING THE RAW DATA ENTERED

DURING THE COURSE OF THE LAB

○ HALLV7.PNG
 FIGURE 1 USED IN THIS LAB REPORT

○ FIGURE2.PNG

 FIGURE 2 USED IN THIS LAB REPORT

○ IMG_1279.JPG
 FIGURE 3 USED IN THIS LAB REPORT

○ IMG_1277.JPG
 FIGURE 5 USED IN THIS LAB REPORT

Page | 27
1 Edwin Hall (1879). "On a New Action of the Magnet on Electric Currents". American Journal
of Mathematics (American Journal of Mathematics, Vol. 2, No. 3) 2 (3): 287–
92. doi:10.2307/2369245. Retrieved 2010-10-30

2 Georgia State University. (n.d.). The Hall Effect. Retrieved October 30, 2010, from
HyperPhysics: http://hyperphysics.phy-astr.gsu.edu/hbase/magnetic/hall.html

3 Qiu, Y. (1997, April 27). Basics of Hall Effect: History. Retrieved October 30, 2010, from
Johns Hopkins University: http://www.pha.jhu.edu/~qiuym/qhe/node1.html

4 See3: Applications

5 Klitzing, K. von; Dorda, G.; Pepper, M. (1980). "New Method for High-Accuracy
Determination of the Fine-Structure Constant Based on Quantized Hall Resistance". Phys.
Rev. Lett. 45 (6): 494–49

6 "Press Release: The 1985 Nobel Prize in Physics". Nobelprize.org. 30 Oct 2010
http://nobelprize.org/nobel_prizes/physics/laureates/1985/press.html

7 Honeywell. (n.d.). Hall Effect Sensing and Application. Retrieved October 30, 2010, from
Honeywell: http://content.honeywell.com/sensing/prodinfo/solidstate/technical/hallbook.pdf

8 Hugh D. Young, R. A. (2007). University Physics. Pearson Addison-Wesley.

9 Latta, D. G. (n.d.). The Hall Effect in Silver Notes. Frostburg: Frostburg State University.

10 Cain, F. (2004, March 1). Record for Furthest Galaxy is Broken Again. Retrieved October
30, 2010, from Universe Today: http://www.universetoday.com/9347/record-for-furthest-
galaxy-is-broken-again/
11 Microstar Laboratories. (2009). Thermocouple Cold Junctions. Retrieved October 30, 2010,
from Microstar Laboratories: http://www.mstarlabs.com/sensors/thermocouple-cold-
junctions.html

S-ar putea să vă placă și