De Broglie Wavelength Equation: Complete Guide to Matter Waves

In 1924, Louis de Broglie proposed a revolutionary idea: particles of matter, like electrons and protons, exhibit wave-like properties. The de Broglie wavelength equation describes the wavelength associated with any moving particle, forming a cornerstone of quantum mechanics. This guide explains the equation, its derivation, and how to calculate matter wave wavelengths.

The De Broglie Wavelength Formula

The de Broglie wavelength equation relates a particle's wavelength to its momentum:

λ = h / p
De Broglie wavelength = Planck's constant ÷ momentum

Since momentum equals mass times velocity (p = mv), this can also be written as:

λ = h / (m × v)
De Broglie wavelength = Planck's constant ÷ (mass × velocity)

Where:

  • λ (lambda) = de Broglie wavelength (meters)
  • h = Planck's constant = 6.62607015 × 10⁻³⁴ J·s (exact value)
  • p = momentum (kg·m/s)
  • m = mass of the particle (kg)
  • v = velocity of the particle (m/s)

The equation shows that the de Broglie wavelength is inversely proportional to both mass and velocity. Heavier or faster particles have shorter wavelengths, while lighter or slower particles have longer wavelengths.

Historical Background: Wave-Particle Duality

The concept of wave-particle duality emerged from a series of discoveries in the early 20th century:

The Photon: Light as Particles

In 1905, Albert Einstein explained the photoelectric effect by proposing that light consists of discrete packets of energy called photons. Each photon has energy E = hf, where h is Planck's constant and f is the frequency. Einstein showed that a photon's momentum is:

p = E/c = hf/c = h/λ

This established that light, traditionally understood as a wave, also behaves as particles.

De Broglie's Hypothesis: Matter as Waves

Louis de Broglie, in his 1924 PhD thesis, proposed the reverse: if waves (light) can behave as particles, then particles (matter) should also behave as waves. He hypothesized that any particle with momentum p has an associated wavelength:

λ = h/p

This bold hypothesis was initially theoretical, but it was soon confirmed experimentally.

Experimental Confirmation

In 1927, Clinton Davisson and Lester Germer at Bell Labs demonstrated electron diffraction—electrons scattering off a nickel crystal produced an interference pattern, just like X-rays (which are waves). This proved that electrons have wave-like properties. Independently, George Paget Thomson observed electron diffraction through thin metal films. Both experiments confirmed de Broglie's hypothesis, and de Broglie received the Nobel Prize in Physics in 1929.

Physical Constants for Calculations

To calculate de Broglie wavelengths, you need these fundamental constants and particle masses:

Planck's Constant

SymbolValueUnits
h6.62607015 × 10⁻³⁴J·s (joule-seconds)
h4.135667696 × 10⁻¹⁵eV·s (electron volt-seconds)
ℏ = h/2π1.054571817 × 10⁻³⁴J·s (reduced Planck constant)

Particle Masses

ParticleMass (kg)Mass (u)
Electron9.1093837 × 10⁻³¹0.000548580
Proton1.6726219 × 10⁻²⁷1.007276
Neutron1.6749275 × 10⁻²⁷1.008665
Alpha particle6.6446573 × 10⁻²⁷4.001506
Muon1.8835316 × 10⁻²⁸0.1134289

Note: 1 u (unified atomic mass unit) = 1.66054 × 10⁻²⁷ kg

Other Useful Constants

ConstantValueUnits
Speed of light (c)299,792,458m/s
Elementary charge (e)1.602176634 × 10⁻¹⁹C (coulombs)
1 electron volt (eV)1.602176634 × 10⁻¹⁹J (joules)

Calculating De Broglie Wavelength: Step-by-Step

Step 1: Identify the Particle and Its Properties

Determine the mass of the particle (in kg) and its velocity (in m/s). For common particles like electrons or protons, use the standard masses above.

Step 2: Calculate Momentum

If velocity is given directly:

p = m × v

If kinetic energy is given instead, use:

p = √(2mE)

where E is the kinetic energy in joules.

Step 3: Apply the De Broglie Equation

λ = h / p = 6.62607015 × 10⁻³⁴ / p

Step 4: Convert to Appropriate Units

For subatomic particles, wavelengths are often extremely small. Convert from meters to more practical units:

  • nanometers (nm): multiply by 10⁹
  • picometers (pm): multiply by 10¹²
  • angstroms (Å): multiply by 10¹⁰

Worked Examples

Example 1: Electron at Moderate Speed

Problem: Calculate the de Broglie wavelength of an electron moving at 1.0 × 10⁶ m/s.

Given:

  • Mass of electron: m = 9.1094 × 10⁻³¹ kg
  • Velocity: v = 1.0 × 10⁶ m/s
  • Planck's constant: h = 6.626 × 10⁻³⁴ J·s

Solution:

  1. Calculate momentum: p = mv = (9.1094 × 10⁻³¹)(1.0 × 10⁶) = 9.1094 × 10⁻²⁵ kg·m/s
  2. Apply de Broglie equation: λ = h/p = (6.626 × 10⁻³⁴)/(9.1094 × 10⁻²⁵)
  3. λ = 7.27 × 10⁻¹⁰ m = 0.727 nm = 7.27 Å

Answer: 0.727 nm (comparable to atomic dimensions)

Example 2: Electron Accelerated Through a Potential

Problem: An electron is accelerated through a potential difference of 100 V. What is its de Broglie wavelength?

Given:

  • Potential difference: V = 100 V
  • Kinetic energy gained: E = eV = (1.602 × 10⁻¹⁹)(100) = 1.602 × 10⁻¹⁷ J

Solution:

  1. Calculate momentum from kinetic energy: p = √(2mE)
  2. p = √(2 × 9.1094 × 10⁻³¹ × 1.602 × 10⁻¹⁷)
  3. p = √(2.919 × 10⁻⁴⁷) = 5.40 × 10⁻²⁴ kg·m/s
  4. λ = h/p = (6.626 × 10⁻³⁴)/(5.40 × 10⁻²⁴) = 1.23 × 10⁻¹⁰ m

Answer: 0.123 nm = 1.23 Å

Shortcut formula for electrons: λ (nm) = 1.226 / √V, where V is in volts

Check: 1.226 / √100 = 1.226 / 10 = 0.1226 nm ✓

Example 3: Thermal Neutrons

Problem: Calculate the de Broglie wavelength of a thermal neutron at room temperature (T = 300 K).

Given:

  • Mass of neutron: m = 1.6749 × 10⁻²⁷ kg
  • Average kinetic energy: E = (3/2)kT, where k = 1.381 × 10⁻²³ J/K

Solution:

  1. Calculate thermal energy: E = (3/2)(1.381 × 10⁻²³)(300) = 6.21 × 10⁻²¹ J
  2. Calculate momentum: p = √(2mE) = √(2 × 1.6749 × 10⁻²⁷ × 6.21 × 10⁻²¹)
  3. p = √(2.08 × 10⁻⁴⁷) = 4.56 × 10⁻²⁴ kg·m/s
  4. λ = h/p = (6.626 × 10⁻³⁴)/(4.56 × 10⁻²⁴) = 1.45 × 10⁻¹⁰ m

Answer: 0.145 nm = 1.45 Å

Thermal neutrons have wavelengths comparable to atomic spacings, making them ideal for neutron diffraction studies of crystal structures.

Example 4: Proton at High Energy

Problem: A proton has kinetic energy of 1 MeV. What is its de Broglie wavelength?

Given:

  • Mass of proton: m = 1.6726 × 10⁻²⁷ kg
  • Kinetic energy: E = 1 MeV = 1 × 10⁶ × 1.602 × 10⁻¹⁹ J = 1.602 × 10⁻¹³ J

Solution:

  1. Calculate momentum: p = √(2mE) = √(2 × 1.6726 × 10⁻²⁷ × 1.602 × 10⁻¹³)
  2. p = √(5.36 × 10⁻⁴⁰) = 2.32 × 10⁻²⁰ kg·m/s
  3. λ = h/p = (6.626 × 10⁻³⁴)/(2.32 × 10⁻²⁰) = 2.86 × 10⁻¹⁴ m

Answer: 28.6 fm (femtometers)

This wavelength is comparable to nuclear dimensions (~1-10 fm), making MeV protons useful for probing nuclear structure.

Example 5: Baseball (Macroscopic Object)

Problem: Calculate the de Broglie wavelength of a 145 g baseball thrown at 40 m/s (about 90 mph).

Given:

  • Mass: m = 0.145 kg
  • Velocity: v = 40 m/s

Solution:

  1. Calculate momentum: p = mv = (0.145)(40) = 5.8 kg·m/s
  2. λ = h/p = (6.626 × 10⁻³⁴)/(5.8) = 1.14 × 10⁻³⁴ m

Answer: 1.14 × 10⁻³⁴ m

This wavelength is about 10⁻²⁰ times smaller than an atomic nucleus! For macroscopic objects, the de Broglie wavelength is so incredibly small that wave-like behavior is completely undetectable. This explains why we don't observe quantum effects in everyday life.

Relativistic De Broglie Wavelength

When particles move at speeds approaching the speed of light, relativistic effects become significant. The relativistic momentum is:

p = γmv = mv / √(1 - v²/c²)
Relativistic momentum, where γ is the Lorentz factor

The de Broglie wavelength then becomes:

λ = h / (γmv) = h√(1 - v²/c²) / (mv)

For particles with known kinetic energy, the relativistic momentum is:

p = √(E² + 2Em₀c²) / c
where E is kinetic energy and m₀ is rest mass

Example: Relativistic Electron

Problem: Calculate the de Broglie wavelength of an electron traveling at 0.8c (80% the speed of light).

Solution:

  1. Calculate the Lorentz factor: γ = 1/√(1 - 0.8²) = 1/√(1 - 0.64) = 1/√0.36 = 1/0.6 = 1.667
  2. Calculate relativistic momentum: p = γmv = 1.667 × 9.109 × 10⁻³¹ × 0.8 × 3 × 10⁸
  3. p = 3.64 × 10⁻²² kg·m/s
  4. λ = h/p = 6.626 × 10⁻³⁴ / 3.64 × 10⁻²² = 1.82 × 10⁻¹² m = 1.82 pm

Answer: 1.82 pm (0.00182 nm)

Compare this to the non-relativistic calculation, which would give λ = h/(mv) = 3.03 pm. The relativistic correction reduces the wavelength by about 40%.

Applications of De Broglie Wavelengths

Electron Microscopy

Electron microscopes exploit the short de Broglie wavelength of accelerated electrons to achieve much higher resolution than optical microscopes. The resolution limit of any imaging system is approximately equal to the wavelength used.

  • Optical microscope: Limited by visible light wavelengths (~500 nm), resolution ~200 nm
  • Scanning Electron Microscope (SEM): Electrons at 10-30 kV, λ ≈ 0.01-0.007 nm, practical resolution ~1-10 nm
  • Transmission Electron Microscope (TEM): Electrons at 100-300 kV, λ ≈ 0.004-0.002 nm, resolution ~0.1-0.2 nm

Modern aberration-corrected TEMs can resolve individual atoms!

Electron Diffraction

When electrons pass through crystalline materials, they diffract according to their de Broglie wavelength. This technique, called electron diffraction, is used to:

  • Determine crystal structures
  • Analyze thin films and surfaces
  • Study phase transitions
  • Characterize nanomaterials

Low-energy electron diffraction (LEED) uses electrons with wavelengths of 0.5-5 Å, ideal for probing surface atomic arrangements.

Neutron Diffraction

Thermal neutrons (E ~ 25 meV, λ ~ 1.8 Å) are particularly useful for studying crystal structures because:

  • They scatter off atomic nuclei, complementing X-ray diffraction which scatters off electron clouds
  • They can locate light atoms (like hydrogen) that are nearly invisible to X-rays
  • They have magnetic moments, allowing study of magnetic structures
  • They penetrate deeply into materials

Atom Interferometry

Even entire atoms can exhibit wave-like behavior. Atom interferometers use the de Broglie wavelength of cold atoms to make extremely precise measurements of:

  • Gravitational acceleration
  • Rotation rates (for navigation)
  • Fundamental constants
  • Gravitational waves (proposed future detectors)

Understanding the Hydrogen Atom

The de Broglie hypothesis helps explain why electrons in atoms occupy discrete energy levels. For a stable orbit, the electron's de Broglie wave must form a standing wave that fits exactly around the orbit circumference:

2πr = nλ = nh/p
n = 1, 2, 3, ... (orbit must contain whole number of wavelengths)

This leads directly to the Bohr quantization condition: L = mvr = nℏ

Comparison: De Broglie Wavelengths of Different Particles

To illustrate how mass and velocity affect the de Broglie wavelength, consider particles at the same kinetic energy (1 eV = 1.602 × 10⁻¹⁹ J):

ParticleMass (kg)Velocity at 1 eV (m/s)λ at 1 eV
Electron9.109 × 10⁻³¹5.93 × 10⁵1.23 nm
Proton1.673 × 10⁻²⁷1.38 × 10⁴28.6 pm
Neutron1.675 × 10⁻²⁷1.38 × 10⁴28.6 pm
Alpha particle6.645 × 10⁻²⁷6.94 × 10³14.3 pm
Uranium atom3.95 × 10⁻²⁵9.0 × 10²1.86 pm

At the same energy, lighter particles have longer wavelengths because they move faster (E = ½mv²), and the wavelength scales as 1/√m for non-relativistic particles at fixed energy.

The Heisenberg Uncertainty Principle Connection

The de Broglie wavelength is intimately connected to the Heisenberg uncertainty principle. If a particle has a well-defined wavelength λ, it has a well-defined momentum p = h/λ. But a wave with a single wavelength extends infinitely in space—its position is completely undefined.

The uncertainty principle states:

Δx · Δp ≥ ℏ/2
Position uncertainty × momentum uncertainty ≥ h/4π

To localize a particle (reduce Δx), you need a superposition of many wavelengths (wave packet), which increases the uncertainty in momentum (and thus wavelength). This fundamental trade-off is a direct consequence of wave-particle duality.

Useful Formulas for Specific Situations

Electrons Accelerated Through Potential V

λ (nm) = 1.226 / √V
V in volts, non-relativistic approximation

Thermal Particles at Temperature T

λ = h / √(3mkT)
Using average thermal kinetic energy E = (3/2)kT

Particle with Kinetic Energy E

λ = h / √(2mE)
Non-relativistic, E in joules

Photon-Equivalent Wavelength

For comparison, a photon with the same energy as a particle's kinetic energy has wavelength:

λ_photon = hc / E

This is always much longer than the de Broglie wavelength of a massive particle with the same energy.

Why Macroscopic Objects Don't Show Wave Behavior

The de Broglie equation applies to all matter, but wave effects are only observable when the wavelength is comparable to relevant physical dimensions (like slits, obstacles, or atomic spacings).

For macroscopic objects:

  • A 70 kg person walking at 1 m/s has λ = 9.5 × 10⁻³⁶ m
  • A dust particle (10⁻⁹ kg) drifting at 1 mm/s has λ = 6.6 × 10⁻²² m
  • A bacterium (10⁻¹² kg) swimming at 10 μm/s has λ = 6.6 × 10⁻¹⁷ m

Even for a bacterium, the wavelength is 10¹⁰ times smaller than a proton! There are no physical structures small enough to diffract such tiny wavelengths, so quantum interference effects are completely unobservable for macroscopic objects. This is why classical mechanics works perfectly well for everyday objects.

The boundary between quantum and classical behavior is an active area of research. Experiments have demonstrated wave-like interference with molecules containing over 2,000 atoms, pushing the limits of where quantum mechanics applies.

Particle Properties Comparison Table

When calculating de Broglie wavelengths, you need accurate values for particle mass, charge, and other properties. The following comprehensive table compares the key physical properties of common particles encountered in quantum mechanics calculations.

ParticleMass (kg)Mass (MeV/c²)Charge (e)SpinClassification
Electron (e⁻)9.1094 × 10⁻³¹0.511-11/2Lepton
Proton (p)1.6726 × 10⁻²⁷938.3+11/2Baryon
Neutron (n)1.6749 × 10⁻²⁷939.601/2Baryon
Muon (μ⁻)1.8835 × 10⁻²⁸105.7-11/2Lepton
Alpha particle (⁴He²⁺)6.6447 × 10⁻²⁷3727.4+20Composite nucleus
Deuteron (²H⁺)3.3436 × 10⁻²⁷1875.6+11Composite nucleus
Pion (π⁺)2.4880 × 10⁻²⁸139.6+10Meson
Tau (τ⁻)3.1675 × 10⁻²⁷1776.9-11/2Lepton
Hydrogen atom (H)1.6735 × 10⁻²⁷938.801/2Atom
Helium-4 atom (⁴He)6.6465 × 10⁻²⁷3728.400Atom
Carbon-12 atom (¹²C)1.9926 × 10⁻²⁶11177.900Atom
Fullerene (C₆₀)1.1970 × 10⁻²⁴~6.7 × 10⁵00Molecule

The mass ratio between particles is particularly important for understanding their de Broglie wavelengths. The proton is 1,836 times heavier than the electron, meaning that at the same velocity, a proton's de Broglie wavelength is 1,836 times shorter. The muon, at about 207 times the electron mass, sits between the electron and proton, and muon diffraction experiments have confirmed de Broglie's hypothesis for this particle as well.

The inclusion of fullerene (C₆₀) in the table is significant. In 1999, researchers at the University of Vienna demonstrated quantum interference with C₆₀ molecules, proving that even molecules containing 60 carbon atoms (with a mass over a million times that of an electron) exhibit wave-particle duality. More recent experiments have shown interference with molecules exceeding 25,000 atomic mass units.

De Broglie Wavelengths at Various Velocities

The de Broglie wavelength depends on both mass and velocity. This table shows how wavelength changes across a wide range of velocities for three important particles: the electron, proton, and neutron. All values use the non-relativistic formula λ = h/(mv), except where noted.

VelocityElectron λProton λNeutron λContext
1 × 10³ m/s727 nm0.396 nm0.396 nmSlow thermal motion
1 × 10⁴ m/s72.7 nm0.0396 nm (39.6 pm)0.0396 nmModerate thermal motion
1 × 10⁵ m/s7.27 nm3.96 pm3.96 pmWarm plasma / accelerated
1 × 10⁶ m/s0.727 nm (7.27 Å)0.396 pm0.396 pmElectron microscopy
1 × 10⁷ m/s0.0727 nm (0.727 Å)0.040 pm0.040 pmHigh-energy accelerator
0.01c (3 × 10⁶ m/s)0.242 nm0.132 pm0.132 pmMildly relativistic
0.1c (3 × 10⁷ m/s)0.0242 nm*0.0132 pm*0.0132 pm*Relativistic corrections needed
0.5c (1.5 × 10⁸ m/s)4.20 pm*2.29 fm*2.28 fm*Relativistic (γ = 1.155)
0.9c (2.7 × 10⁸ m/s)1.18 pm*0.643 fm*0.642 fm*Highly relativistic (γ = 2.294)
0.99c (2.97 × 10⁸ m/s)0.316 pm*0.172 fm*0.172 fm*Ultra-relativistic (γ = 7.089)

*Relativistic values calculated using λ = h/(γmv), where γ = 1/√(1 - v²/c²).

Several key observations emerge from this table. First, electron wavelengths at typical electron microscopy speeds (10⁵-10⁷ m/s) range from about 7 nm down to 0.07 nm, which is why electron microscopes can resolve atomic structures. Second, neutron wavelengths at thermal velocities (corresponding to room temperature, roughly 2,200 m/s) are around 0.18 nm, which is comparable to interatomic spacings in crystals, making thermal neutrons ideal for diffraction studies. Third, the relativistic correction becomes significant above about 0.1c, reducing the wavelength compared to the non-relativistic prediction.

Historical Timeline of Quantum Mechanics Milestones

The de Broglie wavelength equation did not emerge in isolation. It was part of a revolutionary transformation in physics during the early 20th century. The following timeline places de Broglie's contribution in the context of the broader development of quantum mechanics.

YearScientist(s)MilestoneSignificance
1900Max PlanckQuantized energy emission (E = nhf)Introduced Planck's constant h; resolved black-body radiation problem
1905Albert EinsteinPhotoelectric effect and photon conceptShowed light has particle nature; E = hf for individual photons
1913Niels BohrBohr model of the hydrogen atomQuantized electron orbits; explained hydrogen spectral lines
1923Arthur ComptonCompton scatteringConfirmed photon momentum p = h/λ through X-ray scattering
1924Louis de BroglieMatter wave hypothesis (λ = h/p)Proposed all particles have wave-like properties; PhD thesis
1925Werner HeisenbergMatrix mechanicsFirst complete mathematical formulation of quantum mechanics
1926Erwin SchrodingerWave equation (Schrodinger equation)Described matter waves mathematically; equivalent to matrix mechanics
1926Max BornProbabilistic interpretation of wave function|ψ|² gives probability density; statistical nature of QM
1927Davisson and GermerElectron diffraction from nickel crystalExperimental confirmation of de Broglie wavelength for electrons
1927G.P. ThomsonElectron diffraction through thin metal filmIndependent confirmation of electron wave behavior
1927Werner HeisenbergUncertainty principle (Δx·Δp ≥ ℏ/2)Fundamental limit on simultaneous measurement precision
1928Paul DiracRelativistic quantum mechanics (Dirac equation)Unified quantum mechanics with special relativity; predicted antimatter
1929Louis de BroglieNobel Prize in PhysicsAwarded for discovery of the wave nature of electrons
1932James ChadwickDiscovery of the neutronOpened path to neutron diffraction studies
1936Mitchell and PowersFirst neutron diffraction experimentConfirmed de Broglie waves for neutrons
1961Claus JonssonDouble-slit experiment with electronsDemonstrated single-electron interference pattern
1974Merli, Missiroli, PozziSingle-electron double-slit experiment recordedFilmed buildup of interference pattern one electron at a time
1999Arndt et al. (Vienna)C₆₀ fullerene interferenceWave behavior demonstrated for molecule of 720 atomic mass units
2019Fein et al. (Vienna)Interference of molecules >25,000 amuLargest objects to show quantum interference (2,000 atoms)

De Broglie's 1924 hypothesis was a turning point in physics. Before his work, waves and particles were considered entirely separate phenomena. Light was known to behave as both (through Young's double-slit experiment and Einstein's photoelectric effect), but no one had suggested that matter itself might exhibit wave behavior. The experimental confirmations by Davisson-Germer and Thomson within just three years of the hypothesis transformed it from a bold theoretical speculation into an established physical fact.

The progressive demonstration of matter-wave interference with ever-larger objects continues to this day. Each new record pushes the boundary between the quantum and classical worlds, addressing one of the deepest questions in physics: where exactly does quantum behavior give way to classical behavior? The current record holders are organic molecules containing over 2,000 atoms, with de Broglie wavelengths on the order of femtometers.

Summary

The de Broglie wavelength equation λ = h/p = h/(mv) is fundamental to quantum mechanics:

  • Wave-particle duality: All matter exhibits wave-like properties
  • Inverse relationships: Wavelength decreases with increasing mass or velocity
  • Planck's constant (h = 6.626 × 10⁻³⁴ J·s) sets the scale for quantum effects
  • Practical importance: Enables electron microscopy, diffraction studies, and atom interferometry
  • Classical limit: Macroscopic objects have wavelengths far too small to observe wave effects

Use our de Broglie wavelength calculator to quickly compute matter wave wavelengths for electrons, protons, neutrons, and custom particles.

Frequently Asked Questions

The de Broglie wavelength is the wavelength associated with any moving particle, calculated as λ = h/p where h is Planck's constant and p is the particle's momentum. It represents the wave-like nature of matter in quantum mechanics.

Use λ = h/√(2mE), where m is the particle mass and E is the kinetic energy in joules. For electrons accelerated through voltage V, the shortcut is λ (nm) = 1.226/√V.

Macroscopic objects have incredibly small de Broglie wavelengths (10⁻³⁴ m or smaller). There are no structures small enough to diffract such wavelengths, making wave effects completely undetectable. Quantum effects only become visible at atomic and subatomic scales.

Both are calculated from λ = h/p, but photons are massless while matter particles have mass. For the same energy, a photon has a much longer wavelength than a massive particle because its momentum equals E/c, while a particle's momentum is √(2mE).