Fact-checked by Grok 2 weeks ago

Oil drop experiment

The oil drop experiment, devised and conducted by American physicist Robert A. Millikan starting in 1909, is a landmark measurement in physics that precisely determined the of the and provided conclusive evidence for the quantization of electric charge. In the experiment, microscopic oil droplets—roughly one-thousandth of a millimeter in diameter—were sprayed into a horizontal chamber between two large parallel metal plates forming an air condenser, where they became charged through exposure to such as X-rays or light. By observing the droplets' terminal velocities as they fell under gravity alone and then rose or were suspended when a strong (up to 6,000 volts per centimeter) was applied between the plates, Millikan balanced the gravitational and electrostatic forces acting on each droplet to calculate its net charge. This method built on earlier work by J.J. Thomson, who had estimated the electron's charge-to-mass ratio, but Millikan's innovation of using oil droplets instead of allowed for more stable, long-term observations by minimizing and convection currents through thermostatic control of the apparatus. Over hundreds of trials involving thousands of droplets, Millikan found that the charges were always multiples of a base unit, which he quantified as e = 4.774 \times 10^{-10} electrostatic units (equivalent to approximately $1.592 \times 10^{-19} coulombs in modern terms). Published in in , the results not only confirmed the discrete nature of —aligning with Benjamin Franklin's early hypothesis of electricity as indivisible particles—but also enabled the first accurate determination of Avogadro's constant when combined with gas law data. The experiment's significance extends to its role in solidifying the model of , earning Millikan the 1923 for this and related work on photoelectric effects. Despite later revelations that Millikan selectively reported data to align with theoretical predictions, the core findings have withstood scrutiny and remain a foundational of charge quantization in undergraduate physics laboratories worldwide.

Historical Background

Pre-Millikan Attempts to Measure Charge

Prior to Robert Millikan's oil drop experiment in 1909, several physicists, primarily at the under J.J. Thomson, attempted to determine the elementary e following Thomson's 1897 discovery of the . These early efforts focused on measuring the charge carried by ions or charged particles in gases, often using methods involving water droplets formed around ions in humid air. The approaches relied on balancing gravitational and electrostatic forces on falling droplets, similar in to Millikan's later refinement, but suffered from challenges such as droplet evaporation, variable sizes, and imprecise velocity measurements, leading to approximate values with significant uncertainties. One of the first direct attempts was by J.S.E. Townsend in 1897–1898, who developed the falling-drop method using saturated clouds of ionized by X-rays or other means. Townsend measured the mass of individual droplets by observing their under in humid air, then determined the total charge on a droplet by its absorption of ions and subsequent deflection in an . He estimated the charge on gaseous ions to be approximately $1 \times 10^{-19} C, suggesting it represented an atomic unit, though his results varied due to difficulties in controlling droplet uniformity. Building on this, J.J. Thomson conducted measurements in 1898 using X-rays to ionize air in a vessel containing , where ions served as nuclei for droplet . By quantifying the total charge via electrical current and the number of droplets via their sedimentation, Thomson calculated e \approx 6.5 \times 10^{-10} esu, equivalent to about $2.1 \times 10^{-19} C—a value roughly twice the modern accepted figure of $1.602 \times 10^{-19} C. In 1899, Thomson refined the approach with the , liberating electrons from a plate using light and measuring the charge on resulting ions, yielding e \approx 6.8 \times 10^{-10} esu or $2.2 \times 10^{-19} C; this work also confirmed consistency with charges. Harold A. Wilson improved these techniques in 1903 by incorporating an to suspend or control droplet motion more precisely, measuring sedimentation velocities of charged water droplets in air. His method produced a more accurate estimate of e \approx 3.1 \times 10^{-10} esu, or about $1.0 \times 10^{-19} C, closer to the true value but still affected by and measurement errors. These pre-Millikan experiments established the existence of a fundamental charge unit but lacked the precision to demonstrate its quantization clearly, as droplet charges often appeared as multiples without consistent resolution.

Millikan's Motivation and Collaboration

At the turn of the , the discovery of the by J.J. Thomson in 1897 had revolutionized , but the precise value of the electron's charge remained elusive, with Thomson's early estimates of from ion and photoelectric experiments yielding approximate results around 6 × 10^{-10} electrostatic units, while cathode ray work provided the e/m ratio. Robert Millikan, then a at the , was motivated to devise a more accurate method to determine this , aiming to confirm whether was fundamentally particulate and quantized in discrete units, as suggested by earlier theoretical ideas from George Stoney and experimental hints from Thomson. This pursuit was driven by Millikan's broader goal of conducting "crucial experiments" to test the unitary nature of , amid growing evidence from and studies that demanded a definitive value for the charge quantum. In 1908, Millikan, along with Louis Begeman, attempted to measure charges on water droplets ionized in air, building on Harold A. Wilson's 1903 droplet balance method, but evaporation caused inconsistent droplet sizes and unreliable velocity measurements, preventing precise charge determinations. To overcome this, Millikan collaborated with his graduate student starting in December 1909; Fletcher, seeking a topic, proposed replacing with oil droplets of low to minimize evaporation and enable stable observations over extended periods. Fletcher contributed to designing a rudimentary apparatus and analyzing early data, while Millikan refined the setup and performed most observations; their joint efforts yielded preliminary results by May 1910, with the first publication appearing in September of that year, establishing charges as multiples of approximately 4.77 × 10^{-10} electrostatic units. Though Fletcher expected co-authorship on the initial paper, Millikan published it solely under his name, leading to some tension, but Fletcher completed his PhD in 1911 using subsequent data. Later refinements involved assistance from other collaborators, such as Dr. Yoshio Ishida for gas-specific observations, but the core innovation stemmed from the Millikan-Fletcher partnership, which provided the precision needed to solidify the electron charge as a fundamental constant. This work not only resolved prior uncertainties but also earned Millikan the 1923 Nobel Prize in Physics for demonstrating the discrete nature of electric charge.

Theoretical Principles

Forces Acting on Oil Droplets

In the oil drop experiment, charged oil droplets suspended between two horizontal plates experience four primary forces that determine their motion: the gravitational force, the buoyant force, the viscous drag force, and the electrostatic force. These forces must be analyzed to isolate the charge on each droplet, as the terminal velocities of the droplets under different conditions allow for the calculation of the charge magnitude. The gravitational force acts downward on the droplet and is given by F_g = m g, where m is the mass of the oil droplet and g is the . The mass m is determined from the droplet's volume and the density of the oil, typically m = \frac{4}{3} \pi r^3 \rho_{oil}, with r as the radius and \rho_{oil} as the density (approximately 875 kg/m³). This force pulls the droplet toward the lower plate. Opposing the gravitational force is the buoyant force, arising from the displacement of air by the droplet. According to , this upward force is F_b = \frac{4}{3} \pi r^3 \rho_{air} g, where \rho_{air} is the (about 1.2 kg/m³ at standard conditions). Although small compared to the gravitational force—typically reducing the effective weight by less than 1%—it is included for precision in the effective gravitational force: F_{eff} = (m - m_{air}) g, with m_{air} = \frac{4}{3} \pi r^3 \rho_{air}. The viscous drag force resists the droplet's motion through the air and follows Stokes' law for spherical particles at low Reynolds numbers: F_d = 6 \pi \eta r v, where \eta is the viscosity of air (approximately $1.8 \times 10^{-5} Pa·s), r is the radius, and v is the terminal velocity. This force acts opposite to the direction of motion and balances other forces at terminal velocity. For submicron droplets, a Cunningham correction factor is applied to account for slip at the air-droplet interface, modifying the drag to F_d = 6 \pi \eta r v (1 + A e^{B l / r}), where A and B are empirical constants and l is the mean free path of air molecules. However, Millikan's analysis primarily relied on the basic Stokes' form, with corrections introduced later. The electrostatic force is F_e = q E, where q is the charge on the droplet and E is the uniform electric field strength between the plates (E = V / d, with V the applied voltage and d the plate separation). This force can be upward or downward depending on the sign of q and the field direction, but in the experiment, the field is adjusted so that negatively charged droplets (from ) experience an upward force opposing . When the electric field is applied, the droplet reaches a new terminal velocity, enabling the isolation of q from the balance of forces. At without the , the effective gravitational force balances the drag: (m - m_{air}) g = 6 \pi \eta r v_g where v_g is the falling . With the field on, for an upward-moving droplet: q E = (m - m_{air}) g + 6 \pi \eta r v_r where v_r is the rising . Solving these equations yields q = \frac{(m - m_{air}) g (v_g + v_r)}{E v_g}, demonstrating how the forces interplay to measure discrete multiples of the .

Derivation of Charge Measurement Equations

In the oil drop experiment, the charge q on a droplet is determined by balancing the forces acting on it under different conditions: without an and motion under an applied . The key forces are the effective gravitational force (accounting for ), the viscous drag force given by , and the electrostatic force. Millikan assumed that for drag, F_d = 6 \pi \eta r v, holds for the small oil droplets, where \eta is the of air, r is the droplet , and v is the . This assumption was justified by the low for the droplets, ensuring . When no electric field is applied, the droplet reaches a terminal falling velocity v_g downward, where the effective weight balances the drag force: \frac{4}{3} \pi r^3 (\rho - \rho_a) g = 6 \pi \eta r v_g Here, \rho is the density of the oil, \rho_a is the density of air, and g is gravitational acceleration; the left side represents the buoyant weight W. Solving for the radius r: r = \sqrt{ \frac{9 \eta v_g}{2 g (\rho - \rho_a)} } This equation allows r to be calculated from the measured v_g, using known values of \eta, \rho, \rho_a, and g. Millikan calibrated \eta experimentally to account for any deviations from ideal Stokes' law at small scales. With an E = V / d applied (where V is the voltage across plates separated by distance d), the droplet can be made to rise with v_e upward, assuming the field opposes for a positively charged drop (or vice versa for negative). The balance of forces is then: q E = W + 6 \pi \eta r v_e = 6 \pi \eta r (v_g + v_e) Substituting W = 6 \pi \eta r v_g from the no-field case yields the charge: q = \frac{6 \pi \eta r (v_g + v_e)}{E} Inserting the expression for r gives a complete formula in terms of measurable quantities: q = \frac{18 \pi d}{V} \sqrt{ \frac{ \eta^3 v_g (v_g + v_e)^2 }{ 2 g (\rho - \rho_a)} } In practice, Millikan often used cases where the drop fell slowly under the field (v_f) instead of rising, leading to analogous forms like q = \frac{6 \pi \eta r (v_g - v_f)}{E} for field-assisted falling, but the rising case provided clearer measurements for small charges. These derivations enabled precise computation of q for individual drops, revealing quantization in multiples of the elementary charge e.

Experimental Apparatus

Key Components and Design

The oil drop experiment's apparatus, as designed by Robert Millikan in his 1913 study, centered on a closed chamber containing two horizontal parallel plates to generate a . These plates, each approximately 22 cm in diameter, were polished to optical flatness within two wavelengths of sodium to ensure field uniformity, and separated by about 1.6 cm using precisely machined or echelon blocks for accurate spacing. The upper plate featured a small hole through which oil droplets could enter the region between the plates, allowing for controlled observation of their motion under gravity and the applied field. Oil droplets were introduced via a simple atomizer, such as a perfume sprayer filled with refined clock oil, which dispersed fine droplets (typically 1–2 μm in radius) into the chamber above the upper plate; these then fell through the hole into the field region, where they acquired charge through collisions with ionized air molecules. Ionization was achieved using X-rays from a Coolidge tube or emanations from a 500 mg sample of radium bromide placed near the chamber, enabling the droplets to gain or lose electrons and thus exhibit discrete charge values. The chamber itself was constructed from glass or metal to minimize air currents, with provisions for temperature stabilization—such as surrounding water baths and cupric chloride cells—to maintain air viscosity constant within 0.02°C, critical for precise drag force calculations. For observation, the apparatus incorporated a low-power or positioned horizontally to view the droplets' vertical motion through a small window in the chamber wall, with side illumination provided by a to enhance visibility against a dark background. An adjustable high-voltage (up to several kilovolts) connected to the plates allowed Millikan to balance the electric force against , suspending individual or measuring their terminal velocities during rise and fall. This design emphasized simplicity and precision, enabling repeated measurements on thousands of droplets over extended periods (up to hours per ) without significant or contamination.

Oil Droplet Preparation and Ionization

In Millikan's apparatus, oil droplets were generated using a simple atomizer, which produced a fine mist of microscopic particles by forcing air through a containing the . The selected was of high purity and low , such as refined clock or , to minimize and size changes during the experiment; typical droplet radii ranged from 1 to 2 micrometers. This mist was introduced into an upper reservoir chamber above the main observation region, where excess droplets settled or escaped, allowing only a controlled number to proceed. The droplets entered the observation chamber by falling under through a precisely drilled pinhole, approximately 0.05 in , in the upper of two horizontal brass plates separated by about 16 . This pinhole served to isolate individual droplets, preventing overcrowding and enabling clear microscopic observation via a low-power aligned perpendicular to the plates. The chamber itself was enclosed to maintain stable air conditions, with the plates forming a parallel-plate capable of supporting high voltages up to several kilovolts. Neutral oil droplets acquired charge primarily through ionization of the surrounding air, achieved by directing X-rays from a small into the space between the plates. The X-rays, generated by accelerating electrons onto a metal target, produced a of ion pairs in the air, consisting of positive s and free electrons. As uncharged droplets fell through this ionized medium, they captured these charged particles—most often electrons, resulting in a net negative charge—due to the droplets' large surface area relative to the ions. This process allowed for reproducible charging, with the charge magnitude varying based on exposure time and intensity; moreover, repeated enabled adjustment of the charge on a single droplet for multiple measurements. In initial trials, frictional charging during contributed incidentally, but X-ray became the controlled method to ensure variability and precision in charge values.

Measurement Procedure

Droplet Selection and Observation

In Millikan's oil drop experiment, oil droplets were introduced into the apparatus by spraying a fine mist of oil using an ordinary atomizer, producing droplets approximately one thousandth of a millimeter (1 μm) in diameter that entered the space between two horizontal metal plates through a small . The chamber was carefully thermostated to eliminate currents, ensuring the droplets' motion was undisturbed by air movements. Suitable droplets were selected based on their visibility and slow terminal falling speed under alone, typically on the order of 0.01 to 0.05 mm/s, which allowed for precise timing over the observation path and indicated small size and low initial charge for stable tracking. A cloud of droplets was initially observed, but only one isolated droplet was chosen to avoid interactions, using a low-power fitted with a calibrated or micrometer to monitor its position against fixed scale markings. Once selected, the droplet was charged primarily through ionization of the surrounding air by X-rays (or alternatively alpha rays from ), enabling gas ions to attach to the droplet and confer an multiple. The falling motion under was timed as the droplet traversed a 1 cm vertical path in the , yielding the gravitational v_g, after which the (up to 6000 V/cm between 16 mm-spaced, 22 cm-diameter plates) was applied to reverse the motion, and the rising velocity v_r was similarly measured by over the same distance. The field was toggled on and off repeatedly to obtain multiple velocity readings from the same droplet, ensuring consistency and allowing observation periods extending up to hours for reliable data. The observation setup emphasized a dust-free, vibration-isolated environment to maintain the droplet's laminar motion, with illumination adjusted to highlight the droplet against the dark field without heating the air. This methodical selection and repeated observation of individual droplets enabled the isolation of charge effects from other forces acting on the particle.

Field Application and Velocity Measurements

In the Millikan oil drop experiment, the application of the electric field is a critical step following droplet selection, enabling the measurement of the charge on the droplet by balancing gravitational and electrostatic forces. The electric field is generated between two horizontal parallel metal plates separated by 16 mm, with a high-voltage DC power supply providing up to approximately 9,600 V across the plates (22 cm diameter), resulting in a uniform field strength E up to 6,000 V/cm. The polarity of the field is adjusted using a switch to direct the force on negatively charged droplets upward against gravity; for instance, the top plate is set negative relative to the bottom to produce an upward force on the droplet. This field can be turned on or off, or reversed, to control the droplet's motion, allowing it to either rise, fall, or remain suspended if the voltage is precisely tuned to balance the forces. Velocity measurements begin with the droplet falling freely under gravity in the absence of the electric field, where it reaches a terminal velocity v_f due to the balance between gravitational force and viscous drag from air. The experimenter observes the droplet through an illuminated microscope equipped with a reticle scale in the eyepiece, marking fixed intervals. A stopwatch is used to time the droplet's passage between these marks, typically over multiple trials to ensure precision, yielding the average fall velocity v_f = \Delta x / t_f, where \Delta x is the distance (1 cm path) and t_f the average time. Once v_f is recorded, the electric field is activated, causing the charged droplet to rise against gravity to a new terminal velocity v_r, determined similarly by timing its upward motion over the same reticle intervals, with v_r = \Delta x / t_r. These rise and fall velocities are essential for subsequent charge calculations, as they allow determination of the droplet's radius and mass via Stokes' law before computing the charge from the force balance q = \frac{m g (v_f + v_r)}{v_f E}, where m is the droplet mass and g is gravitational acceleration. Measurements are repeated after ionizing the droplet to alter its charge, capturing multiple data points for the same droplet to track charge quantization.

Data Analysis and Results

Calculating Droplet Charges

In the oil drop experiment, the charge q on a selected droplet is determined by measuring its terminal velocities under two conditions: without an applied and directed motion with the field applied. The terminal velocity in , denoted v_g, allows calculation of the droplet's radius a, while the upward terminal velocity v_e under the E provides the data needed to compute q. This method relies on balancing the gravitational force, viscous drag, and electric force acting on the droplet at terminal velocities, assuming steady-state motion where is zero. The radius a is first derived from the free-fall measurement. The effective gravitational force on the droplet is m' g, where m' = \frac{4}{3} \pi a^3 (\rho - \rho_{\text{air}}) is the effective mass accounting for buoyancy, \rho is the oil density, and \rho_{\text{air}} is the air density. At terminal velocity v_g, this balances the viscous drag force given by Stokes' law: m' g = 6 \pi \eta a v_g, where \eta is the viscosity of air. Substituting the expression for m' yields the quadratic relation: \frac{4}{3} \pi a^3 (\rho - \rho_{\text{air}}) g = 6 \pi \eta a v_g Simplifying gives: a^2 = \frac{9 \eta v_g}{2 (\rho - \rho_{\text{air}}) g} Thus, a = \sqrt{\frac{9 \eta v_g}{2 (\rho - \rho_{\text{air}}) g}} This equation assumes the droplet is spherical and that Stokes' law applies without correction for small Reynolds numbers; Millikan introduced a empirical correction factor (1 + \frac{b}{a}) to the drag force for droplets smaller than about 1 μm, where b is a constant derived from experimental calibration. Velocities v_g and v_e are obtained by timing the droplet's traversal of a known distance between microscope crosshairs. With the radius known, the charge q is calculated from the motion under the applied . Assuming the droplet carries a negative charge (as ionized by X-rays), the downward-directed field causes an upward electric force q E. When the field is on and the droplet rises at v_e, the forces balance as: electric force upward equals gravitational force downward plus drag force downward (opposing the upward motion): q E = m' g + 6 \pi \eta a v_e Since m' g = 6 \pi \eta a v_g from the free-fall case, substitution yields: q E = 6 \pi \eta a v_g + 6 \pi \eta a v_e = 6 \pi \eta a (v_g + v_e) Solving for q: q = \frac{6 \pi \eta a (v_g + v_e)}{E} Here, E = V / d, with V the applied voltage and d the plate separation. Including the Stokes' correction, the drag terms become $6 \pi \eta a (1 + b/a) v, so the charge formula adjusts to: q = \frac{6 \pi \eta a (1 + b/a) (v_g + v_e)}{E} This process is repeated for multiple droplets and multiple ionizations per droplet to accumulate charge values, ensuring statistical reliability. Millikan reported charges typically in the range of integer multiples of approximately $4.77 \times 10^{-10} esu (equivalent to $1.59 \times 10^{-19} C), demonstrating quantization.

Quantization and Determination of Elementary Charge

The charges calculated for the oil droplets in Millikan's experiment revealed a striking : every measured charge q was an integer multiple of a fundamental unit e, the , with the multiplier n typically ranging from 1 to several hundred. This observation, based on data from over 100 droplets, provided the first direct experimental evidence for the quantization of , confirming that is not continuous but in nature. No fractional multiples were found, even within the experimental uncertainties, underscoring the nature of charge carriers like the . To determine the value of [e](/page/E!), Millikan focused on a subset of droplets for which the charge remained over multiple measurements, minimizing systematic errors from charge variation. He computed each [q](/page/Q) using the balance of forces and then sought the common that rendered all [q](/page/Q) / [e](/page/E!) as integers. By grouping the charges into series (e.g., assuming the smallest were [e](/page/E!), next $2[e](/page/E!), etc.) and averaging the implied values of [e](/page/E!) from [q](/page/Q) / n, weighted by the of each , he arrived at [e](/page/E!) = 4.774 \times 10^{-10} esu (electrostatic units). This corresponds to $1.592 \times 10^{-19} C in units, with an estimated of 0.2%. The involved iterative refinement to ensure across the , effectively using the distribution of charges to isolate the elementary unit. This determination not only quantified e but also validated theoretical predictions from electrolysis and other indirect methods, bridging atomic theory with precise measurement. Millikan's approach highlighted the importance of statistical analysis in handling measurement errors, as small deviations in individual q values were accommodated by the integer-multiplicity assumption.

Comparisons and Accuracy

Millikan's Reported Value

In his seminal 1913 paper detailing the oil drop experiment, Robert Millikan reported the value of the elementary electrical charge as e = (4.774 \pm 0.009) \times 10^{-10} electrostatic units (esu). This determination stemmed from analysis of selected measurements on 58 oil droplets out of 107 measured between February and April 1912 (as part of work spanning 1909–1913), where the charges observed on the drops were found to be integer multiples of this fundamental unit. Millikan emphasized that the value represented an average derived from balancing gravitational, viscous, and electric forces on the droplets, with careful corrections for factors such as air density and droplet . To contextualize the precision, Millikan noted an uncertainty of \pm 0.009 \times 10^{-10} esu. In his later 1917 summary in The Electron, he refined the estimate slightly to e = (4.774 \pm 0.002) \times 10^{-10} esu, incorporating additional data but upholding the core result from the oil drop method. This esu value is equivalent to approximately $1.592 \times 10^{-19} coulombs in modern units, using the conversion factor of $1 esu = 3.33564 \times 10^{-10} C. Using this charge alongside the electrochemical , Millikan also computed Avogadro's number as N = 6.062 \times 10^{23} molecules per mole, providing one of the earliest precise links between atomic-scale charge quantization and macroscopic electrochemical measurements. These results solidified the quantization of electric charge and earned Millikan the 1923 for his work on the .

Alignment with Modern Measurements

Millikan's final reported for the , derived from measurements on 58 oil droplets over 60 days, was e = 4.774 \times 10^{-10} esu, equivalent to approximately $1.592 \times 10^{-19} C. This result demonstrated charge quantization with high precision for the era, establishing e as the fundamental unit from which all observed charges were multiples. Millikan's was later found slightly low due to an error in air determination, contributing to the deviation from modern measurements. The modern accepted value, fixed exactly since the 2019 SI redefinition, is e = 1.602176634 \times 10^{-19} C. Millikan's measurement deviates by about 0.61%, a remarkably close agreement given the experimental challenges, including accurate determination of oil viscosity and gravitational effects without contemporary instrumentation. This alignment validates the oil drop method's foundational accuracy, as subsequent refinements in technique—such as improved ionization control and video microscopy in lab replications—yield values within 1% of the CODATA standard, often achieving uncertainties below 0.5%. Contemporary measurements of e rely primarily on methods like the and Josephson junctions, which provide uncertainties at the parts-per-billion level and confirm the quantization observed by Millikan without relying on mechanical droplet dynamics. Despite these advances, the oil drop experiment's result remains a for early 20th-century precision, underscoring its role in bridging classical and . Modern adaptations, including searches for fractional charges (e.g., signatures at \pm e/3), build directly on Millikan's framework but detect no deviations from integer multiples of e.

Controversies and Criticisms

Data Selection Practices

In Robert Millikan's 1913 publication detailing the oil drop experiment, he reported measurements from 58 oil droplets, asserting that these represented "all of the drops experimented upon during 60 consecutive days" without any selection bias. However, analysis of his laboratory notebooks reveals that he recorded data from approximately 100 to 140 drops over 63 days between late 1911 and April 1912, excluding around 40 to 75 observations before publication. These exclusions were justified in the notebooks by notes on factors such as unstable droplet motion, suspected contamination, or inconsistencies in the electric field application, which Millikan deemed unreliable for accurate charge calculations. Millikan's selection criteria emphasized droplets that exhibited consistent terminal velocities under gravitational and , allowing for precise determination of charge by balancing the gravitational and electrostatic forces on each droplet. He discarded drops showing anomalous charges, such as those suggesting fractions of the (e.g., a drop on April 16, 1912, with a charge approximately two-thirds of expected multiples), as these could indicate experimental errors rather than subelectron entities. This practice was partly motivated by Millikan's presupposition of charge quantization in discrete elementary units, leading him to interpret outliers as artifacts to be excluded in order to refine the average value of e. The data selection practices ignited controversy, particularly with physicist Felix Ehrenhaft, who conducted similar experiments and reported subelectron charges from his own droplet observations, accusing Millikan of cherry-picking to suppress evidence of charge continuity. Ehrenhaft's claims, published in the early , prompted Millikan to intensify scrutiny of his data, excluding drops that might align with subelectron hypotheses to bolster his quantization argument. Despite this, statistical reanalyses of Millikan's full notebook data by later historians, such as Allan Franklin, indicate that including the discarded drops would alter the statistical but not significantly shift the value of e from Millikan's reported 4.77 × 10^{-10} esu. Historiographical assessments of these practices remain divided. Scholars like Gerald Holton and Allan Franklin argue that Millikan's selections reflected legitimate scientific judgment amid experimental uncertainties, such as variable air and droplet evaporation, rather than deliberate fraud, noting that similar discretionary choices were common in early 20th-century physics. Conversely, critics highlight the discrepancy between Millikan's published claim of non-selection and the evident filtering as a form of methodological , potentially misleading readers and violating contemporary standards for transparent reporting in responsible conduct of . David Goodstein, analyzing the notebooks, concludes that while the practice was not intentionally deceptive—given its minimal impact on results—it would today require explicit justification to avoid accusations of data manipulation. Overall, the episode underscores the tension between presupposition-driven experimentation and objective data handling in foundational scientific work.

Role of Psychological and Methodological Biases

In Robert Millikan's oil drop experiment, psychological biases, particularly , played a significant role in shaping data interpretation and selection. Millikan approached the measurements with a strong presupposition of charge quantization in discrete units, influenced by his atomistic worldview and visualizations of electrons as tangible corpuscles. This led him to favor observations aligning with an value near 4.77 × 10^{-10} esu, while dismissing anomalies that suggested subelectron charges, as seen in his rejection of data from Ehrenhaft's competing experiments. Historian Holton noted that Millikan's notebooks reveal a tendency to rate observation sets with stars based on their fit to expected results, effectively weighting data to confirm his rather than exploring alternatives. Methodological biases compounded these psychological influences through selective data practices. Millikan recorded approximately 140 runs but reported only 58 in his paper, excluding drops due to perceived irregularities such as incomplete observations or deviations from preferred speed ranges (10–40 seconds per fall), which he justified as minimizing errors from or chronograph inaccuracies. This selection on dependent variables—choosing data that reduced variance and aligned with preconceived values—violated principles of unbiased estimation later emphasized in modern methodological frameworks like the credibility revolution. For instance, he discarded an observation yielding a charge 30% lower than his final value, attributing it to without duplication, despite its potential validity. These biases were not unique to Millikan but reflected broader effects in early 20th-century physics, where theory-laden observations influenced experimental and . In the Millikan-Ehrenhaft dispute, Millikan's insistence on unitary charges led him to interpret Ehrenhaft's smaller charge measurements as experimental artifacts rather than evidence of subelectrons, a stance driven by the prevailing of discrete electron charges. Richard Jennings analyzed Millikan's notebooks and concluded that while the selection was a form of scientific judgment amid controversy, it misrepresented the comprehensiveness of his , raising ethical concerns under contemporary responsible conduct standards without constituting outright . Ultimately, these practices contributed to Millikan's reported value being slightly lower than modern measurements, highlighting how intertwined psychological and methodological biases can skew foundational results.

Legacy and Modern Perspectives

Impact on Atomic and Quantum Physics

The oil drop experiment, conducted by Millikan in 1909–1913, yielded the first accurate direct measurement of the elementary , determining its value as e = 1.592 \times 10^{-19} C with an uncertainty of just 0.2%. This result unequivocally demonstrated that is quantized, occurring only in multiples of this fundamental unit, thereby confirming the as an indivisible particle with discrete charge rather than a continuous fluid-like entity as some earlier theories suggested. By refuting claims of subelectron charges proposed by Ehrenhaft and establishing charge quantization experimentally, the work provided critical for the particulate nature of electricity, a cornerstone of early 20th-century . This precise value of e was instrumental in advancing theoretical models of the atom. In Niels Bohr's 1913 quantum model of the , the charge features prominently in the foundational equations; for instance, the , which governs frequencies, is given by R = \frac{m_e e^4}{8 \epsilon_0^2 h^3 c}, where e^4 determines the scale of atomic energies and radii. Bohr explicitly acknowledged the significance of Millikan's , stating that key relations in his model held "to within the considerable accuracy with which, especially through the beautiful investigations of Millikan, the charge of the is known." Without such a reliable e, quantitative predictions of Bohr's quantized orbits and energy levels—essential for explaining hydrogen's —would have lacked the precision needed to gain traction among physicists. Beyond the Bohr model, Millikan's findings accelerated the broader acceptance of quantum theory by underscoring the discrete, non-classical behavior of subatomic entities. The experiment's confirmation of charge quantization paralleled emerging ideas of energy quantization in blackbody radiation and the photoelectric effect, reinforcing the shift from classical to quantum paradigms and facilitating developments in quantum electrodynamics and atomic spectroscopy. This foundational constant enabled subsequent calculations of atomic properties, such as ionization energies and fine structure, profoundly shaping our understanding of quantum interactions within atoms.

Recent Historiographical Reassessments

In the late 1970s, physicist and historian Gerald Holton provided a seminal historiographical analysis of Millikan's oil drop experiment by examining the researcher's laboratory notebooks, revealing that Millikan selectively included only about 42% of his measurements in published results, excluding data that did not align with his presupposition of charge quantization. Holton's work framed this selection not as outright but as influenced by theoretical commitments, including the atomistic view of electricity prevalent at the time, and highlighted the broader Millikan-Ehrenhaft dispute, where Felix Ehrenhaft's similar experiments yielded evidence for subelectron charges, challenging Millikan's conclusions. Building on Holton, of Millikan's unpublished data argued that most exclusions were technically justified due to experimental inconsistencies, such as unreliable droplet behavior, though a small portion appeared motivated by the need to refute Ehrenhaft's fractional charge claims; Franklin concluded that including all data would have slightly increased the statistical error but not altered the quantization finding or the value of the . This perspective shifted the toward viewing Millikan's practices as standard for the era, emphasizing methodological rigor amid experimental challenges like air currents and droplet evaporation. Mansoor Niaz's 2005 appraisal further reassessed the controversy, synthesizing Holton and Franklin to argue that closure on the debate is possible only by acknowledging the role of presuppositions in data interpretation; Niaz noted that Millikan and Ehrenhaft obtained comparable results but diverged due to differing theoretical frameworks—Millikan's commitment to discrete charges versus Ehrenhaft's continuity—illustrating how scientific consensus emerges through community adjudication rather than pure empiricism. In a 2009 rational reconstruction, Niaz extended this to critique chemistry textbooks for omitting the Ehrenhaft rivalry and presenting the experiment inductivistly, as a straightforward empirical triumph, thereby underrepresenting the interpretive complexities and perseverance required in science. More recent scholarship, such as David H. Glass's 2023 examination, reinforces these themes by underscoring the non-decisive role of alone in scientific decisions, citing Holton's notebook analysis to show how Millikan's atomistic presuppositions shaped exclusions, while Franklin's demonstrates the robustness of the core result despite such influences. Overall, post-2000 portrays the oil drop experiment as a in the interplay of , , and , moving beyond accusations of to illuminate the human elements of scientific progress.

References

  1. [1]
    [PDF] Robert A. Millikan - Nobel Lecture
    This is why the body chosen to replace the pith ball was an individual oil- droplet about a thousandth of a millimeter in diameter blown out of an ordinary ...
  2. [2]
    Determination of the Charge on an Electron - chemteam.info
    Millikan started his work on electron charge in 1906 and continued for seven years. His 1913 article announcing the determination of the electron's charge is a ...
  3. [3]
    Sir John Sealy Townsend - Britannica
    In 1897 Townsend developed the falling-drop method for measuring e. His procedure, which used saturated clouds of water, was later improved and ultimately led ...Missing: 1897-1903 | Show results with:1897-1903
  4. [4]
  5. [5]
  6. [6]
  7. [7]
    Remembering the oil-drop experiment - Physics Today
    May 1, 2007 · In 1910 Robert Millikan of the University of Chicago announced that he and Harvey Fletcher had developed a successful means of isolating and ...
  8. [8]
    [PDF] My work with Millikan on the oil-drop experiment
    Harvey Fletcher (1884-1981) directed acous- tical and, later, physical research at Bell Labo- ratories from 1925 to 1952, developing hear- ing aids and ...Missing: collaboration | Show results with:collaboration
  9. [9]
    Robert A. Millikan – Facts - NobelPrize.org
    By measuring how the various drops of oil moved about, Millikan showed that their charge always was a multiple of a precisely determined charge—the electron's ...
  10. [10]
    [PDF] The Millikan Oil-Drop Experiment - University of Toronto
    This experiment first described by [Millikan, 1913] is based on the fact that different forces act on an electrically charged oil drop moving in the homogeneous ...
  11. [11]
    [PDF] MILLIKAN OIL DROP EXPERIMENT - De Anza College
    In this experiment an electrical force of varying magnitude is introduced to change the motion of the falling drop by an ionization source. By measuring the ...
  12. [12]
    On the Elementary Electrical Charge and the Avogadro Constant
    The Millikan oil drop experiment, published in final form in 1913, demonstrated that charge comes in discrete chunks and was a bridge between classical ...
  13. [13]
  14. [14]
    Landmarks—Millikan Measures the Electron’s Charge
    ### Summary of Droplet Preparation and Ionization in Millikan's Experiment
  15. [15]
    Millikan's oil drop experiment (Theory) : Modern Physics Virtual Lab
    Now the total force acting on drop is Fe-F'v-FG=0. F'V is the new viscous force under the action of electric field. Millikan repeated the experiment no.
  16. [16]
    Millikan Oil Drop - UCSB Physics
    May 2, 2025 · Typically, a whole cloud of droplets of various sizes and charges is introduced into the viewing chamber by squeezing the bulb of an atomizer.
  17. [17]
    [PDF] Millikan Oil Drop Data Analysis: - UCI Physics and Astronomy
    The experiment consists of raising a tiny, electrically charged oil drop in an electric field and then lowering it again. To raise it you apply a constant ...<|control11|><|separator|>
  18. [18]
    On the Elementary Electrical Charge and the Avogadro Constant
    Published August 1, 1913 | Version public. Journal Article Open. On the Elementary Electrical Charge and the Avogadro Constant. Creators. Millikan, R. A.. Style.
  19. [19]
  20. [20]
    Robert A. Millikan and the Oil Drop Experiment - AIP Publishing
    Oct 1, 2019 · Millikan determined the charge of the electron to be 4.77 ± 0.009×10−10 electrostatic units (1.592×10−19 coulombs).
  21. [21]
    <i>Landmarks</i>—Millikan Measures the Electron's Charge
    Jan 20, 2012 · The Millikan oil drop experiment, published in final form in 1913, demonstrated that charge comes in discrete chunks and was a bridge between ...
  22. [22]
    Introduction to the constants for nonexperts 19001920
    Millikan's final value, reported in 1917, was: (4.774 ± 0.002) x 10-10 esu (esu being the electrostatic unit, one of the units of charge in the centimeter ...
  23. [23]
    August, 1913: Robert Millikan Reports His Oil Drop Results
    Aug 1, 2006 · Robert Millikan's famous oil drop experiment, reported in August 1913, elegantly measured the fundamental unit of electric charge.
  24. [24]
    elementary charge - CODATA Value
    elementary charge $e$. Numerical value, 1.602 176 634 x 10-19 C. Standard uncertainty, (exact). Relative standard uncertainty, (exact).Missing: 2018 | Show results with:2018
  25. [25]
    [PDF] CODATA RECOMMENDED VALUES OF THE FUNDAMENTAL ...
    CODATA RECOMMENDED VALUES OF THE FUNDAMENTAL PHYSICAL CONSTANTS: 2018 ... elementary charge e. 1.602 176 634 × 10−19 (exact). C proton magnetic moment.
  26. [26]
    [PDF] Millikan oil drop experiment - NYU Physics department
    Robert Millikan reported in 1910 a method for measuring the charge e of a single electron by monitoring the motion of a droplet of oil suspended in a chamber ...
  27. [27]
    [PDF] On the Elementary Electrical Charge and the Avogadro Constant
    ON THE ELEMENTARY ELECTRICAL CHARGE AND THE. AVOGADRO CONSTANT. BV R. A. MILLIKAN. I. INTRODUCTOR Y. HE experiments herewith reported were undertaken ...
  28. [28]
    Data selection and responsible conduct: Was Millikan a fraud?
    Aug 6, 2025 · This paper discusses two main issues that arise in assessing his conduct, whether he was intentionally misleading and whether he actually did ...Missing: scholarly | Show results with:scholarly
  29. [29]
    The Millikan Case - Discrimination Versus Manipulation of Data
    Jan 1, 2000 · An examination of Millikan's own papers and notebooks reveals that he picked and chose among his drops. That is, he exercised discrimination ...Missing: sources | Show results with:sources
  30. [30]
    An appraisal of the controversial nature of the oil drop experiment
    A closure to the controversy is possible if we recognize that Millikan's data selection procedure depended primarily on his commitment to his presuppositions.Missing: practices sources
  31. [31]
  32. [32]
    [PDF] Niels Bohr - Nobel Lecture
    This relation has been shown to hold to within the con- siderable accuracy with which, especially through the beautiful investiga- tions of Millikan, the ...
  33. [33]
    Robert Millikan - Magnet Academy
    ### Summary of Millikan's Impact on Bohr's Quantum Theory and Photoelectric Effect
  34. [34]
    (PDF) Milliken's Oil Drop Experiment and the Role of Empirical ...
    Oct 15, 2023 · ... 1913 he arrived at a result of 4.77. units using "All of the [58] drops experimented on during 60 consecutive days" (Holton, 1978, p.63;.
  35. [35]
    An Appraisal of the Controversial Nature of the Oil Drop Experiment
    Aug 8, 2025 · A more recent appraisal by Niaz concluded that if Millikan's handling of data is to be understood with no reference to his presuppositions, then ...Missing: reassessment | Show results with:reassessment
  36. [36]
    The Oil Drop Experiment: A Rational Reconstruction of the Millikan ...
    Aug 9, 2025 · (a) The Millikan-Ehrenhaft controversy can open a new window for students, demonstrating how two well-trained scientists can interpret the same ...