AP Physics 1
AP Physics 1: Algebra-Based is an introductory, algebra-based college-level physics course offered by the College Board as part of its Advanced Placement program, designed to cultivate students' understanding of physics through inquiry-based investigations of core concepts in Newtonian mechanics and related topics.[1] It serves as the equivalent of a first-semester university physics course without calculus, focusing on developing scientific practices such as creating models, using mathematical routines, and engaging in argumentation from evidence.[2] The course is structured around seven big ideas: systems, fields, force interactions, change, conservation laws, waves, and probabilistic nature of quantum measurements (though the latter is introductory in this context).[2] Key topics are organized into eight units, including kinematics (10%–15% exam weighting), forces and translational dynamics (18%–23%), work, energy, and power (18%–23%), linear momentum (10%–15%), torque and rotational dynamics (10%–15%), energy and momentum of rotating systems (5%–8%), oscillations (5%–8%), and fluids (10%–15%).[1] At least 25% of instructional time is devoted to hands-on laboratory work, with a minimum of eight inquiry-based labs to mirror a semester-long college physics laboratory experience.[2] Prerequisites include completion of geometry, concurrent enrollment in Algebra II or equivalent, and familiarity with high school algebra and basic trigonometry; prior physics exposure is recommended but not required.[2] The AP Physics 1 exam, administered annually in May, is a three-hour hybrid digital assessment with Section I: 40 multiple-choice questions (80 minutes, 50% of the score) and Section II: four free-response questions (100 minutes, 50%), testing both conceptual understanding and problem-solving skills.[3] Scores of 3 or higher on the exam may qualify students for college credit or advanced placement at over 3,300 institutions worldwide, though policies vary by school.[2]History
Development and Introduction
The Advanced Placement (AP) program, administered by the College Board, began offering physics courses in the 1955-1956 school year as part of its initial slate of ten subjects, with the curriculum emphasizing foundational topics in mechanics and electricity to align with introductory college-level study.[4][5] These early courses aimed to provide high school students with rigorous preparation equivalent to first-year university physics, focusing on core principles without advanced mathematics. Over the decades, the program evolved to meet changing educational needs, but by the 1990s, the algebra-based offering had expanded into AP Physics B, a comprehensive one-year survey covering mechanics, electricity and magnetism, waves, optics, thermodynamics, and modern physics.[6] By the early 2010s, concerns arose among educators and researchers that AP Physics B's broad scope compromised depth and rigor, leading to superficial coverage of topics and misalignment with the investigative approaches used in college introductory physics courses.[7] In response, the College Board collaborated with university professors, high school teachers, and the National Research Council to redesign the curriculum, announcing in October 2012 the replacement of AP Physics B with two separate algebra-based courses: AP Physics 1, centered on mechanics and fluids, and AP Physics 2, addressing thermodynamics, electricity and magnetism, optics, and modern physics (revised in 2024 to move the fluids unit to AP Physics 1).[7][8][2] This split allowed for greater focus and sequencing, enabling students to build skills progressively over two years. AP Physics 1 launched in the fall of 2014 for the 2014-2015 school year, with the first exam administered in May 2015 to 171,074 students.[9] The redesign's core rationale was to foster deeper conceptual understanding, hands-on laboratory skills, and inquiry-based learning rather than rote memorization, aligning more closely with evidence-based practices in college physics education and emphasizing science practices such as modeling, data analysis, and experimental design.[10][2] This approach aimed to better prepare students for scientific reasoning and real-world applications, addressing longstanding critiques of the prior course's breadth-over-depth structure.[8]Major Revisions
Since its launch in 2014 as an algebra-based introductory physics course, AP Physics 1 has seen several key revisions to better align with college-level expectations and address pedagogical challenges.[8] The 2019 update to the Course and Exam Description (CED) refined the science practices from six original categories to three main ones—creating representations, mathematical routines, and scientific questioning and argumentation—while placing greater emphasis on modeling physical systems and experimental design to foster deeper conceptual understanding.[11][12] This revision aimed to streamline skill development, integrating more guided inquiry labs to support student engagement with real-world applications.[12] A substantial content change took effect for the 2024-2025 school year with the addition of Unit 8: Fluids, transferred from AP Physics 2 to provide a more comprehensive foundation in mechanics. This unit covers density, pressure, buoyancy, and fluid dynamics, allocated 10-15% of the exam weighting to encourage exploration of how fluids interact with forces and motion.[2] Concurrently, the exam transitioned to a hybrid digital format in 2025—where multiple-choice questions are completed in the Bluebook app and free-response answers are handwritten—and remains hybrid as of 2025, enhancing security through features like randomized question sets and improving accessibility with accommodations like text-to-speech.[3][13] For the 2025-2026 school year, minor clarifications refined wording in the CED for consistency across AP Physics courses, such as splitting free-response question parts for clearer scaffolding and updating the front matter to emphasize career preparation alongside college readiness.[11] These ongoing adjustments respond to performance trends, with scores of 3 or higher historically ranging from about 40-50% (e.g., 46.4% in 2024), though the 2025 pass rate rose to 67.3%, potentially due to curriculum enhancements and exam format changes.[14][15][11]Course Overview
Objectives and Prerequisites
AP Physics 1 aims to cultivate students' understanding of physics principles through inquiry-based exploration, modeling of physical systems, application of mathematical routines, and design of experiments, preparing them for introductory college-level algebra-based physics courses.[2] The primary educational goals include fostering scientific inquiry and problem-solving skills, enabling students to connect fundamental concepts like kinematics, dynamics, and energy to real-world phenomena while emphasizing critical thinking over rote memorization.[2] This approach suits students interested in STEM fields who possess solid algebraic foundations but are not yet prepared for calculus-based physics.[2] Key skills developed in the course encompass creating visual representations such as graphs and diagrams, deriving algebraic expressions, estimating numerical quantities, and justifying claims using evidence from experiments or data analysis.[2] At least 25% of instructional time is dedicated to hands-on laboratory activities, where students engage in guided and open-inquiry experiments to apply these skills and deepen conceptual understanding.[2] These elements build foundational abilities that support the subsequent curriculum units.[2] Prerequisites for AP Physics 1 include completion of high school Geometry and concurrent enrollment in Algebra II or an equivalent course, with no prior physics experience required.[2] Students should also have basic familiarity with trigonometry, particularly sine and cosine functions for vector components, which can be reinforced during the course if needed; additionally, comfort with scientific notation and unit conversions is recommended to handle quantitative aspects effectively.[2] This preparation ensures students can focus on physics concepts without being hindered by mathematical barriers, promoting an emphasis on analytical reasoning and experimental design.[2]Structure and Assessment
The AP Physics 1 course is structured as a full academic year program, typically encompassing 140–200 instructional hours depending on the school schedule, such as 45-minute classes meeting five days a week or 90-minute block periods over a semester.[2] It is organized into eight units that build progressively on core concepts, with suggested class periods allocated as follows: Unit 1 (12–17 periods), Unit 2 (22–27 periods), Unit 3 (22–27 periods), Unit 4 (10–15 periods), Unit 5 (15–20 periods), Unit 6 (8–14 periods), Unit 7 (5–10 periods), and Unit 8 (12–17 periods).[2] This pacing allows flexibility for teachers to adapt to local needs while ensuring coverage of essential topics through a balanced sequence of instruction and practice.[2] A key component of the course structure is the requirement for at least 25% of instructional time—approximately 45 minutes per week in a standard five-day schedule—to be devoted to hands-on or virtual laboratory activities.[2] These labs emphasize inquiry-based investigations, where students engage in data collection, graphical analysis, error evaluation, and experimental design, often progressing from guided to open-ended formats.[2] Labs can incorporate low-cost materials or digital simulations to accommodate varying resources, focusing on developing skills in scientific practices such as modeling physical phenomena and interpreting uncertainties.[2] Classroom assessment in AP Physics 1 centers on formative evaluations to monitor student progress and reinforce learning, without mandated weightings that teachers must follow.[2] Per unit, educators typically administer progress checks via AP Classroom, consisting of about 18 multiple-choice questions and 4 free-response questions that align with the course's science practices and unit objectives.[2] Additional assessments include teacher-scored laboratory reports, investigative projects, and problem-solving tasks, which provide ongoing feedback and help students apply concepts in contextual scenarios.[2] The instructional approach prioritizes inquiry-based learning, encouraging students to design and conduct their own investigations to explore physics principles actively.[2] Resources such as AP Classroom's unit guides, question banks, and progress checks support this method by offering scaffolded practice and immediate feedback, while tools like the AP Physics 1 Student Workbook and TIPERs (Tasks Inspired by Physics Education Research) facilitate collaborative and conceptual activities.[2] To ensure accessibility, the course framework promotes accommodations for diverse learners, including English language learners and students with disabilities, through adaptable lab formats, differentiated instruction, and alignment with College Board equity policies that address barriers in underserved communities.[2]Curriculum
Unit 1: Kinematics
Unit 1 of the AP Physics 1 curriculum introduces the foundational concepts of kinematics, which describe the motion of objects in one and two dimensions without considering the forces that cause the motion.[2] This unit emphasizes the use of algebraic representations to model position, velocity, and acceleration, fostering students' ability to analyze and predict motion through inquiry-based investigations.[2] Kinematics serves as the starting point for understanding physical systems and how they change over time, aligning with the course's big ideas of systems (Big Idea 1), interactions and change (Big Idea 3), and conservation laws in changing systems (Big Idea 4).[2] Central to kinematics are the distinctions between scalar and vector quantities. Scalars possess only magnitude, such as distance (the total path length traveled) and speed (the rate of change of distance).[2] Vectors, in contrast, have both magnitude and direction, including displacement (the straight-line change in position from initial to final point), velocity (the rate of change of displacement), and acceleration (the rate of change of velocity).[2] Position refers to an object's location relative to a chosen reference frame, often represented as a vector \vec{r}. Displacement \Delta \vec{x} is then \vec{x} - \vec{x_0}, where \vec{x_0} is the initial position. Average velocity is defined as \vec{v}_{avg} = \frac{\Delta \vec{x}}{\Delta t}, while instantaneous velocity \vec{v} is the limit of this as \Delta t approaches zero, giving the velocity at a specific instant. Similarly, average acceleration \vec{a}_{avg} = \frac{\Delta \vec{v}}{\Delta t}, and instantaneous acceleration \vec{a} is its limit as \Delta t approaches zero.[2] In one-dimensional motion along a straight line, these concepts are applied using the kinematic equations for constant acceleration. The position as a function of time is given by x = x_0 + v_0 t + \frac{1}{2} a t^2 where x_0 is the initial position, v_0 is the initial velocity, a is the constant acceleration, and t is time. The velocity as a function of time is v = v_0 + a t, and the velocity squared as a function of position is v^2 = v_0^2 + 2a (x - x_0).[2] These equations assume uniform acceleration and are derived from the definitions of velocity and acceleration, allowing prediction of motion without calculus. Graphical analysis complements these equations: on a position-time graph, the slope at any point equals the instantaneous velocity; on a velocity-time graph, the slope equals acceleration, and the area under the curve equals displacement; on an acceleration-time graph, the area under the curve equals the change in velocity.[2] Motion in two dimensions builds on one-dimensional principles by treating horizontal and vertical components independently, particularly in projectile motion under constant gravitational acceleration. In such cases, the horizontal motion has zero acceleration (a_x = 0), so v_x = v_{x0} (constant velocity), while vertical motion follows a_y = -g (where g \approx 10 \, \mathrm{m/s^2}), using the same kinematic equations for the y-component. This independence allows the trajectory to be analyzed as the vector sum of these components, resulting in a parabolic path for projectiles launched at an angle.[2] For uniform circular motion, the speed remains constant, but velocity changes direction continuously, implying centripetal acceleration toward the center, though the magnitude of velocity (speed) is uniform.[2] Reference frames and relative motion are also addressed, where velocity relative to different observers varies, but absolute motion is described in an inertial frame.[2] Students represent kinematic motion through multiple formats to deepen understanding. Motion diagrams depict an object's positions at equal time intervals with velocity vectors, showing changes in speed and direction. Verbal descriptions outline scenarios qualitatively, while mathematical equations provide quantitative predictions. Free-body sketches are introduced here as simple diagrams of objects without forces, focusing on position and motion paths.[2] Laboratory activities reinforce these concepts, such as using motion detectors to generate position-time graphs for a cart's constant velocity or acceleration down an incline, analyzing free-fall drops with yo-yos or balls to verify g, or video analysis of projectile launches to separate horizontal and vertical components. For example, rolling balls on tracks demonstrates constant acceleration, allowing students to plot velocity-time graphs and calculate a from the slope.[2] This unit constitutes 10%–15% of the AP Physics 1 exam, testing students' ability to apply kinematic models to describe and predict motion in various representations.[2] These kinematic descriptions of acceleration lay the groundwork for understanding the forces that cause such changes in motion in subsequent units.[2]Unit 2: Force and Translational Dynamics
Unit 2 of the AP Physics 1 curriculum delves into the principles governing translational dynamics, focusing on how forces cause changes in the motion of objects. Building briefly on the kinematic equations from Unit 1 to describe motion, this unit shifts emphasis to the causal role of forces through Newton's laws. Students analyze interactions between objects and their environments, constructing free-body diagrams to represent all forces acting on a system and applying quantitative methods to predict accelerations and equilibrium conditions. The content aligns with enduring understandings that forces characterize interactions and alter momentum, as outlined in the official course framework.[2] Central to the unit are Newton's three laws of motion, which provide the foundational framework for dynamics. Newton's first law, the law of inertia, states that an object remains at rest or in uniform motion along a straight line unless acted upon by a net external force. This principle, originally formulated in Latin as "Corpus omne perseverare in statu suo quiescendi vel movendi uniformiter in directum, nisi quatenus illud vi impressa cogitur statum suum mutare" in Newton's 1687 Philosophiæ Naturalis Principia Mathematica, explains why objects in equilibrium experience zero net force.[16] In AP Physics 1, it is applied to scenarios where multiple forces balance, such as a block at rest on a horizontal surface.[2] Newton's second law quantifies the relationship between net force, mass, and acceleration: \vec{F}_{net} = m \vec{a}, where \vec{F}_{net} is the vector sum of all external forces, m is the object's mass, and \vec{a} is its acceleration. Newton derived this in the same 1687 work, emphasizing that the change in motion is proportional to the motive force impressed and occurs along the line of action.[16] Students use this law to solve for unknown accelerations or forces, distinguishing between equilibrium (\vec{a} = 0, so \vec{F}_{net} = 0) and non-equilibrium cases where net force produces acceleration. Free-body diagrams are essential tools here, isolating an object and diagramming all forces as vectors at a point, with scales and directions accurately represented.[2] Newton's third law asserts that for every action, there is an equal and opposite reaction: if object A exerts a force on object B, then B exerts an equal-magnitude, opposite-direction force on A. This bidirectional interaction, detailed in Newton's Principia as forces between two bodies being equal and oppositely directed along the line joining their centers, applies to all force pairs, such as contact or field forces.[16] In the curriculum, it distinguishes internal forces (which cancel in pairs within a system) from external forces (which determine the system's overall motion).[2] The unit examines key contact and field forces relevant to translational motion. The gravitational force near Earth's surface is given by F_g = mg, where m is mass and g \approx 9.8 \, \mathrm{m/s^2} is the local acceleration due to gravity; this approximation derives from Newton's universal law of gravitation for weak fields, as presented in the Principia.[16] The normal force N acts perpendicular to a surface, balancing components of other forces to prevent penetration, such as supporting an object against gravity on a table. Tension T is the pulling force along a massless, inextensible string or rope, always directed away from the object along the string's line. Frictional forces oppose relative motion or tendencies to slide. Static friction f_s satisfies f_s \leq \mu_s N, where \mu_s is the coefficient of static friction and N is the normal force; it adjusts up to a maximum to prevent motion. Kinetic friction, once sliding occurs, is f_k = \mu_k N, constant and opposing the direction of velocity. These empirical relations originated with Guillaume Amontons in 1699, who found friction proportional to normal load and independent of contact area, later refined experimentally by Charles-Augustin de Coulomb in his 1781 memoirs on statics and friction.[17] Spring forces follow Hooke's law, \vec{F} = -k \vec{x}, where k is the spring constant and \vec{x} is the displacement from equilibrium; this linear restoring force was first proposed by Robert Hooke in his 1678 publication De potentia restitutiva, encoded initially as the anagram "ceiiinosssttuv" and revealed as "ut tensio, sic vis" (as the extension, so the force). In AP contexts, these forces are analyzed in isolation or combination using vector addition in free-body diagrams.[2] Applications integrate these concepts to model real-world scenarios. On an inclined plane, the gravitational component parallel to the plane is mg \sin \theta, opposed by friction and resolved to find net force and acceleration down the slope. Atwood machines involve two masses connected by a string over a pulley, where tension and gravity yield system acceleration a = \frac{(m_1 - m_2)g}{m_1 + m_2} for m_1 > m_2, demonstrating third-law pairs and second-law application. Elevator problems explore apparent weight: in an accelerating elevator, the normal force differs from mg, such as N = m(g + a) upward, highlighting equilibrium in non-inertial reference frames from the ground perspective. These examples emphasize solving for unknowns by summing forces in multiple directions, contrasting static equilibrium (net force zero) with dynamic acceleration.[2] For systems of particles, internal forces between components cancel by Newton's third law, so the center of mass accelerates as if all external forces act on the total mass: \vec{F}_{net, ext} = M \vec{a}_{cm}, where M is total mass. This simplifies analysis of composite objects, like connected blocks, by focusing on external interactions. The unit's exam weighting is 18%–23%, testing abilities to create representations, perform calculations, and justify claims using these principles.[2] Laboratory investigations reinforce these topics, such as using force sensors to measure tension, normal, or spring forces in real time, verifying Hooke's law through force-displacement graphs. Friction experiments, like pulling blocks with varying surfaces, determine \mu_s and \mu_k by comparing maximum static force to N or measuring kinetic force constants, often via inclined plane methods to isolate variables. These hands-on activities develop skills in data collection and error analysis, aligning with science practices in the curriculum.[2]Unit 3: Work, Energy, and Power
Unit 3 of AP Physics 1 focuses on the principles of work, energy, and power, shifting the analysis of mechanical systems from the force-based methods of Unit 2 to an energy perspective that emphasizes conservation and transfer.[2] This unit explores how energy is stored, transformed, and transferred in systems, enabling students to solve problems involving motion without directly applying Newton's laws in every step.[2] Key concepts include the work-energy theorem, forms of mechanical energy, and the rate of energy transfer known as power, with applications to everyday scenarios like ramps and springs.[2] The unit typically spans 22–27 class periods and constitutes 18–23% of the multiple-choice section on the AP exam.[2] Work is defined as the transfer of energy that occurs when a force is applied over a displacement, quantified by the scalar product of the force vector and the displacement vector.[2] Mathematically, the work done by a constant force is given byW = \vec{F} \cdot \Delta \vec{r} = F d \cos \theta,
where F is the magnitude of the force, d is the magnitude of the displacement, and \theta is the angle between the force and displacement vectors.[2] For variable forces or non-perpendicular components, work can be calculated by integrating the force over the path or summing components, often referencing force types like gravity or tension from prior units.[2] The net work done on an object equals the change in its kinetic energy, as stated by the work-energy theorem:
W_{\text{net}} = \Delta K = K_f - K_i,
where kinetic energy K = \frac{1}{2} m v^2, with m as mass and v as speed.[2] This theorem allows prediction of final speeds or distances without resolving all forces individually.[2] Mechanical energy in a system consists of kinetic energy and potential energy, with the latter depending on position in a conservative field.[2] Translational kinetic energy is expressed as
K = \frac{1}{2} m v^2,
representing the energy of motion for an object of mass m moving at speed v.[2] Gravitational potential energy near Earth's surface is
U_g = m g h
or more generally U_g = m g y, where g is the acceleration due to gravity, h or y is the height relative to a reference level, and changes are \Delta U_g = m g \Delta h.[2] For elastic systems like springs, potential energy is stored as
U_s = \frac{1}{2} k x^2,
with k as the spring constant and x as the displacement from equilibrium.[2] In isolated systems with only conservative forces, total mechanical energy E = K + U is conserved, so E_i = E_f, allowing energy bar charts or calculations to track transformations between kinetic and potential forms.[2] When nonconservative forces like friction act, mechanical energy is not conserved, and the work done by these forces accounts for the change: W_{nc} = \Delta E = E_f - E_i.[2] For example, on a ramp with friction, the frictional work W_f = -f_k d \cos \phi (where f_k is kinetic friction and \phi is the angle of displacement relative to friction) dissipates energy as heat, reducing the final kinetic or potential energy compared to the frictionless case.[2] Applications include analyzing a block sliding down an inclined plane, where initial potential energy converts to kinetic energy minus frictional losses, or a spring launching a mass up a ramp, balancing elastic, gravitational, and dissipative effects.[2] Power measures the rate at which work is done or energy is transferred, providing insight into the time dependence of mechanical processes.[2] Average power is
P_{\text{avg}} = \frac{W}{t},
the total work divided by the time interval.[2] Instantaneous power is the dot product of force and velocity:
P = \vec{F} \cdot \vec{v} = F v \cos \theta,
useful for varying forces, such as a person lifting an object at constant speed where P = F v if aligned.[2] Efficiency, defined as \eta = \frac{W_{\text{out}}}{W_{\text{in}}} or \eta = \frac{P_{\text{out}}}{P_{\text{in}}}, quantifies how effectively a system converts input energy to useful output, often less than 100% due to nonconservative losses like friction or air resistance.[2] In applications, such as a spring-block system on a ramp, power calculations help determine the time to reach maximum height under frictional conditions.[2] Hands-on labs reinforce these concepts through inquiry-based experiments, comprising about 25% of instructional time.[2] Typical investigations include a spring-loaded ball launcher to verify elastic potential energy conversion to kinetic energy, a cart rolling down a ramp to demonstrate gravitational potential energy conservation, and a block on a frictional ramp to measure nonconservative work and efficiency.[2] These activities emphasize data collection, graphical analysis of energy vs. position, and justification of conservation principles using evidence from motion sensors or force probes.[2]
Unit 4: Linear Momentum
Unit 4 of the AP Physics 1 curriculum introduces linear momentum as a fundamental vector quantity that describes the motion of objects and systems, emphasizing its conservation in isolated interactions.[2] Linear momentum \vec{p} for a single object is defined as the product of its mass m and velocity \vec{v}, expressed as \vec{p} = m \vec{v}, where the direction of \vec{p} aligns with \vec{v}.[2] This definition extends to systems of multiple objects, where the total momentum is the vector sum of individual momenta.[2] The unit builds on prior concepts by applying momentum to analyze changes in motion due to forces over time, contrasting with the scalar nature of energy covered in Unit 3.[2] Impulse represents the effect of a net force acting over a time interval, quantifying the change in linear momentum.[2] Mathematically, impulse \vec{J} is given by \vec{J} = \vec{F}_{\text{net}} \Delta t, where \vec{F}_{\text{net}} is the net force and \Delta t is the time interval; this equals the change in momentum \Delta \vec{p} = m (\vec{v}_f - \vec{v}_i).[2] The impulse-momentum theorem links Newton's second law to momentum changes, allowing predictions of final velocities after interactions like impacts.[2] For example, in vehicle safety, crumple zones extend \Delta t to reduce \vec{F}_{\text{net}}, minimizing \Delta \vec{p} per unit time.[2] In isolated systems—those with no net external force—the total linear momentum remains conserved in all directions, a principle derived from Newton's third law.[2] This conservation holds for the vector sum: \sum \vec{p}_{\text{initial}} = \sum \vec{p}_{\text{final}}.[2] For systems experiencing external impulses, the total momentum changes by the net external impulse.[2] Closed systems approximate isolation if external forces are negligible, while open systems account for external influences like friction.[2] The center-of-mass velocity \vec{v}_{\text{cm}} of a system, defined as \vec{v}_{\text{cm}} = \frac{\sum m_i \vec{v}_i}{\sum m_i}, remains constant without net external force, providing a useful reference for analyzing motion.[2] Collisions between objects illustrate momentum conservation, categorized by kinetic energy behavior from Unit 3.[2] In elastic collisions, both total momentum and kinetic energy are conserved, allowing reversal of velocities in equal-mass one-dimensional (1D) cases.[2] Inelastic collisions conserve momentum but not kinetic energy, with some energy converted to other forms like heat or deformation.[2] Perfectly inelastic collisions, where objects stick together post-impact, maximize kinetic energy loss while conserving momentum, yielding a common final velocity \vec{v}_f = \frac{m_1 \vec{v}_{1i} + m_2 \vec{v}_{2i}}{m_1 + m_2} for two objects in 1D.[2] These principles extend to two-dimensional (2D) collisions, requiring separate conservation equations for x- and y-components.[2] Applications of linear momentum include explosions and recoils, where internal forces cause equal-and-opposite momentum changes, conserving total momentum in isolation.[2] For instance, in a recoil scenario, a cannon's backward momentum equals the forward momentum of the fired cannonball, determined by mass ratios.[2] In 1D collisions, such as two carts on a track, initial velocities predict post-collision outcomes via conservation.[2] 2D collisions, like a puck glancing off another, require vector decomposition to solve for unknowns.[2] These concepts apply to real-world scenarios, such as ballistic projectiles where momentum conservation at impact informs velocity calculations.[2] This unit comprises 10–15% of the AP Physics 1 exam, testing abilities to create diagrams, perform calculations, and justify claims using conservation laws and the impulse-momentum theorem.[2] Laboratory investigations reinforce these ideas through hands-on experiments, such as cart collisions on tracks to measure velocities before and after elastic or inelastic impacts using motion sensors.[18] Air track setups simulate low-friction 1D collisions, verifying momentum conservation quantitatively.[18] Ballistic pendulum labs demonstrate perfectly inelastic collisions, where a projectile embeds in a pendulum bob, combining momentum conservation with energy analysis for height measurements.[18] Spring-loaded "explosion" setups with carts further illustrate conservation in diverging motions.[18]Unit 5: Torque and Rotational Dynamics
Unit 5 of the AP Physics 1 curriculum extends the principles of translational dynamics to rotational motion, focusing on how forces produce torques that cause angular acceleration in rigid bodies. Students learn to describe and analyze rotation using angular quantities analogous to linear ones, such as angular displacement \theta (measured in radians), angular velocity \omega = \frac{\Delta \theta}{\Delta t}, and angular acceleration \alpha = \frac{\Delta \omega}{\Delta t}.[2] The rotational kinematic equations mirror those for linear motion, including \omega = \omega_0 + \alpha t, \theta = \theta_0 + \omega_0 t + \frac{1}{2} \alpha t^2, \omega^2 = \omega_0^2 + 2 \alpha (\theta - \theta_0), and \theta = \theta_0 + \frac{(\omega + \omega_0)}{2} t, allowing predictions of rotational motion under constant angular acceleration.[2] Torque, denoted \vec{\tau}, is the rotational equivalent of force and is defined as the cross product \vec{\tau} = \vec{r} \times \vec{F}, with magnitude \tau = r F \sin \theta, where r is the distance from the axis of rotation to the point of force application, F is the force magnitude, and \theta is the angle between \vec{r} and \vec{F}.[2] The direction of torque follows the right-hand rule, pointing along the axis of rotation.[2] Torque causes changes in angular motion, and linear forces from earlier units serve as components that generate torque when offset from the rotation axis.[2] Rotational inertia, or moment of inertia I, quantifies a body's resistance to angular acceleration and depends on the mass distribution relative to the axis of rotation.[2] For a point mass, I = m r^2, where m is the mass and r is the perpendicular distance from the axis.[2] For rigid objects, I = \sum m_i r_i^2, summing over all mass elements; values are minimized when the axis passes through the center of mass.[2] The parallel axis theorem extends this: I = I_{cm} + M d^2, where I_{cm} is the moment about the center of mass, M is total mass, and d is the distance between parallel axes.[2] Newton's first law for rotation states that an object at rest or rotating with constant angular velocity remains so unless acted upon by a net torque, requiring \sum \vec{\tau} = 0 for rotational equilibrium.[2] In static equilibrium, both net force and net torque must be zero.[2] Newton's second law in rotational form is \vec{\tau}_{net} = I \vec{\alpha}, linking net torque to angular acceleration, analogous to F_{net} = m a.[2] This equation enables analysis of how distributed torques accelerate rigid bodies.[2] Applications of these concepts include analyzing seesaws for balance, where torques from weights on either side must sum to zero for equilibrium; ladders leaning against walls, requiring consideration of friction and normal forces' torques; and pulley systems, where a falling mass generates torque to accelerate a rotating disk.[2] Levers demonstrate mechanical advantage, as torque multiplication allows smaller input forces to produce larger output torques over greater distances, following the principle \tau_{in} = \tau_{out} in equilibrium.[2] This unit comprises 10–15% of the AP Physics 1 exam and typically involves laboratory investigations, such as measuring rotational motion using a falling mass attached to a rotating object to verify \tau = I \alpha.[2][19]Unit 6: Energy and Momentum of Rotating Systems
Unit 6 of the AP Physics 1 curriculum extends the principles of energy and momentum to rotational systems, building on linear analogs to analyze how objects rotate and interact under conservation laws.[2] Students explore rotational kinetic energy and angular momentum, applying these concepts to isolated systems where no external torques or non-conservative forces act, leading to conserved quantities that simplify problem-solving.[2] This unit emphasizes conceptual connections, such as how rotational motion parallels translational dynamics, and includes practical scenarios like spinning objects and rolling bodies.[2] It accounts for 5–8% of the exam weighting in the multiple-choice section.[2] Rotational kinetic energy quantifies the energy associated with an object's rotation about an axis and is given by the formula K_r = \frac{1}{2} I \omega^2, where I is the moment of inertia and \omega is the angular velocity.[2] In systems where only conservative forces do work, mechanical energy is conserved, allowing the sum of translational and rotational kinetic energies plus potential energy to remain constant.[2] For example, a yo-yo released from rest converts gravitational potential energy into both translational kinetic energy of its center of mass and rotational kinetic energy as it unwinds, demonstrating conservation through analysis of its motion.[2] Torque, introduced in Unit 5, relates to changes in rotational energy via work done, where the work-energy theorem for rotation states that net torque times angular displacement equals the change in rotational kinetic energy.[2] Angular momentum for a rigid body rotating about a fixed axis is defined as \vec{L} = I \vec{\omega}, a vector quantity pointing along the axis of rotation according to the right-hand rule.[2] The net torque on the system equals the rate of change of angular momentum, \vec{\tau} = \frac{d \vec{L}}{dt}, which in isolated systems with no external torque leads to conservation of angular momentum.[2] Angular impulse, the product of torque and time or equivalently the change in angular momentum, applies to scenarios where brief interactions alter rotation.[2] This conservation principle explains phenomena in spinning objects, such as a figure skater pulling in their arms to increase spin speed, where the decrease in moment of inertia compensates to keep L constant.[2] Applications of these concepts include gyroscopes, which maintain stable orientation due to conservation of angular momentum; an applied torque causes precession rather than tipping, as the torque vector changes the direction of \vec{L} perpendicularly.[20] In rotational collisions, such as a disk impacting and sticking to a stationary rod pivoted at one end, conservation of angular momentum about the pivot allows calculation of the post-collision angular velocity, treating the system as isolated if friction is negligible.[21] For orbiting satellites, angular momentum conservation about the central body governs stable circular motion, where gravitational torque is zero, maintaining constant L = m v r for the satellite's velocity v and orbital radius r.[2] Combined translation and rotation occur in rolling without slipping, where the linear velocity v of the center of mass relates to angular velocity by v = r \omega, with r as the radius, ensuring the point of contact has zero instantaneous velocity relative to the ground.[2] The total kinetic energy is then K = \frac{1}{2} m v^2 + \frac{1}{2} I \omega^2, which for a solid sphere rolling down an incline from conservation of energy yields an acceleration less than that of pure sliding due to the rotational contribution.[2] Demonstrations, such as racing a ring, disk, and solid cylinder down an incline, illustrate how distribution of mass affects moment of inertia and thus speed, with the solid cylinder arriving first.[2] Laboratory activities reinforce these ideas, such as using a rotating platform to investigate conservation of angular momentum by observing changes in rotation rate when mass distribution varies, akin to the fidget spinner demonstration where the spin axis remains fixed.[2] Another lab involves a rolling disk on a ramp to verify energy conservation and the no-slipping condition through measurements of velocity and angular speed. These experiments emphasize science practices like creating graphical representations of data and justifying claims with evidence from rotational dynamics.[2]Unit 7: Oscillations
Unit 7 of the AP Physics 1 curriculum examines oscillatory motion, with a primary focus on simple harmonic motion (SHM), where objects undergo periodic back-and-forth movement about an equilibrium position. This unit builds on prior concepts such as forces and energy to analyze systems like mass-spring oscillators and pendulums, emphasizing the conditions under which motion approximates SHM. Students learn to model these systems mathematically and graphically, understanding how restoring forces drive the oscillation. The unit accounts for 5–8% of the AP Physics 1 exam.[2] Simple harmonic motion arises when the net restoring force on an object is directly proportional to its displacement from equilibrium and directed opposite to the displacement, expressed as F = -kx, where k is the force constant and x is the displacement. For a mass-spring system, the period of oscillation is given by T = 2\pi \sqrt{\frac{m}{k}}, independent of amplitude for small displacements, where m is the mass. In a simple pendulum, consisting of a mass suspended from a massless string of length L, the period is T = 2\pi \sqrt{\frac{L}{g}}, where g is the acceleration due to gravity; this approximation holds for small angular displacements less than about 15 degrees. A physical pendulum, such as a rigid body pivoting about a fixed axis not through its center of mass, has a period T = 2\pi \sqrt{\frac{I}{mgd}}, where I is the moment of inertia about the pivot, m is the mass, and d is the distance from the pivot to the center of mass.[2] The kinematics of SHM can be described using sinusoidal functions, with position as x = A \cos(\omega t + \phi), where A is the amplitude, \omega = \frac{2\pi}{T} is the angular frequency, t is time, and \phi is the phase constant. Velocity follows as v = -A \omega \sin(\omega t + \phi), and acceleration as a = -\omega^2 x, reaching maximum values of v_{\max} = A \omega and a_{\max} = A \omega^2 at equilibrium and maximum displacement, respectively. Energy in SHM is conserved, with total mechanical energy E = \frac{1}{2} k A^2 constant, partitioned between kinetic energy K = \frac{1}{2} m v^2 and elastic potential energy U = \frac{1}{2} k x^2, as derived from principles of energy conservation introduced in Unit 3. Graphical analysis of SHM involves plotting position, velocity, and acceleration versus time or displacement, revealing phase relationships such as velocity being 90 degrees out of phase with position.[2] Real oscillatory systems often exhibit damped oscillations due to dissipative forces like friction or air resistance, which reduce the amplitude over time while the period remains approximately unchanged for light damping. In driven oscillations, an external periodic force can lead to resonance when the driving frequency matches the natural frequency, causing large amplitude increases, though damping limits the growth to prevent infinite amplitude. Applications of these concepts include analyzing spring-mass systems in shock absorbers and pendulum clocks, where graphical methods help verify sinusoidal behavior and predict motion. Laboratory investigations typically involve timing oscillations in mass-spring setups to determine periods and graphing data to model SHM, reinforcing experimental skills in measurement and analysis.[2]Unit 8: Fluids
Unit 8 of the AP Physics 1 curriculum, newly incorporated in the revised course framework effective for the 2024-25 school year, examines the properties, static behavior, and dynamics of fluids, integrating principles of force, energy, and conservation laws from earlier units. Fluids, defined as materials that flow and conform to container shapes—such as liquids and gases—exhibit behaviors distinct from rigid solids, enabling analyses of pressure distribution and motion under gravity. This unit emphasizes conceptual understanding through algebraic modeling and applies these ideas to real-world scenarios like submerged objects and flowing liquids. It accounts for 10–15% of the multiple-choice questions on the AP Physics 1 exam.[2] Fundamental properties of fluids include density and pressure. Density \rho quantifies mass distribution and is calculated as \rho = \frac{m}{V}, where m is mass and V is volume; for water at standard conditions, \rho \approx 1000 kg/m³, establishing a baseline for comparisons in buoyancy and hydrostatics. Pressure P represents force distributed over area, given by P = \frac{F}{A}, and acts uniformly in all directions within a static fluid due to molecular collisions. In hydrostatic equilibrium, where fluids are at rest, pressure at a depth h below the surface is P = P_0 + \rho g h, with P_0 as the surface pressure (often atmospheric) and \rho g as the gravitational force per unit volume; this relation arises from balancing the weight of the overlying fluid column against pressure differences. Buoyancy explains why objects float or sink in fluids, governed by Archimedes' principle: the upward buoyant force F_b on a submerged or partially submerged object equals the weight of the displaced fluid, F_b = \rho_f V_{\text{disp}} g, where \rho_f is the fluid density and V_{\text{disp}} is the displaced volume.[22] For fully submerged objects, if the object's density exceeds \rho_f, it sinks as F_b < mg; conversely, floating occurs when the object partially displaces fluid such that F_b = mg, with submerged volume fraction \frac{V_{\text{sub}}}{V} = \frac{\rho_{\text{obj}}}{\rho_f}. This principle, derived from integrating hydrostatic pressure over the object's surface, highlights equilibrium under gravitational forces.[22] Fluid dynamics addresses motion in ideal fluids—assumed incompressible, inviscid, and undergoing steady laminar flow—using conservation laws. The continuity equation ensures mass conservation by maintaining constant volume flow rate through varying cross-sections: A_1 v_1 = A_2 v_2, where A is cross-sectional area and v is speed; thus, narrowing pipes accelerate flow to preserve throughput.[23] Bernoulli's equation extends energy conservation along a streamline: P + \rho g h + \frac{1}{2} \rho v^2 = \text{constant}, relating pressure, gravitational potential, and kinetic energy per unit volume; derived from work-energy principles for steady flow, it predicts trade-offs like increased speed reducing pressure.[24] Practical applications demonstrate these concepts in everyday systems. Hydraulics leverages Pascal's principle, where pressure in an enclosed incompressible fluid transmits undiminished in all directions, enabling force amplification in devices like car jacks: \frac{F_2}{A_2} = \frac{F_1}{A_1}, so larger pistons yield greater output forces from smaller inputs.[25] Siphons exploit Bernoulli's equation to transfer liquids over barriers, where flow initiates if the outlet is below the surface level, driven by hydrostatic pressure differences despite temporary height increases. The Venturi effect, a constriction-induced pressure drop, follows from continuity and Bernoulli: as v rises in a narrowed section, P falls, facilitating applications like carburetors or atomizers. Real fluids deviate from ideality due to viscosity, a measure of internal friction that slows flow and promotes laminar-to-turbulent transitions; qualitative effects include energy dissipation, contrasting ideal assumptions.[24] Laboratory investigations reinforce these topics through hands-on inquiry. Students measure buoyancy by immersing objects and recording displaced volumes to verify Archimedes' principle and calculate densities, often using water displacement in graduated cylinders. Flow rate experiments, such as timing efflux from orifices or varying pipe diameters, confirm the continuity equation and apply Bernoulli's equation to predict speeds, for instance, via Torricelli's theorem v = \sqrt{2gh} for fluid exiting a tank hole. These activities, typically spanning 12–17 class periods, develop skills in data collection, graphical analysis (e.g., Bernoulli bar charts), and error evaluation.[2]Exam
Format and Administration
The AP Physics 1 exam is structured as a 3-hour assessment divided into two main sections, each contributing 50% to the total score. Section I consists of 40 multiple-choice questions, to be completed in 80 minutes, and includes both single-select and multi-select question types that may be presented as discrete items or in sets with stimuli or data representations. Section II features 4 free-response questions over 100 minutes, designed to evaluate deeper application of concepts through specific formats: one question on mathematical routines and modeling worth 10 points, one on translation between representations worth 12 points, one on experimental design and analysis worth 10 points, and one on qualitative-quantitative translation worth 8 points.[2] Beginning in 2024, the exam transitioned to a digital format administered via the College Board's Bluebook app, with a hybrid approach implemented for the 2025 administration: multiple-choice questions are completed digitally, while free-response questions are viewed in the app but answered by hand in provided booklets that are subsequently scanned for scoring. Calculators, including four-function, scientific, or graphing models, are permitted for the entire exam duration to support numerical computations and graphical analysis. A digital equation sheet, accessible within Bluebook, supplies all required formulas, physical constants, and conventions—such as those for kinematics, dynamics, and fluids—eliminating the need for students to memorize or reference external materials.[3][26][2] The exam is typically administered during the AP testing window in May, with the 2025 AP Physics 1 exam held on Friday, May 16, at 8:00 a.m. local time; late testing options are available for qualifying students during the week of May 19–23. It is proctored in supervised settings at high schools or authorized testing centers to ensure standardized conditions. Students requiring accommodations, such as extended time or assistive technology, must apply through the College Board's Services for Students with Disabilities (SSD) program, with approvals managed by school coordinators via SSD Online. The exam's structure aligns with the AP Physics 1 curriculum units, testing foundational concepts across kinematics through fluids.[27][28][2]Content and Science Practices
The AP Physics 1 exam evaluates students' mastery of foundational physics concepts through eight units, with multiple-choice questions distributed according to the following approximate weightings: Unit 1 (Kinematics): 10–15%; Unit 2 (Force and Translational Dynamics): 18–23%; Unit 3 (Work, Energy, and Power): 18–23%; Unit 4 (Linear Momentum): 10–15%; Unit 5 (Torque and Rotational Dynamics): 10–15%; Unit 6 (Energy and Momentum of Rotating Systems): 5–8%; Unit 7 (Oscillations): 5–8%; and Unit 8 (Fluids): 10–15%.[2] These units collectively address five big ideas central to classical mechanics: Systems, which explore properties like mass and internal structure; Fields, which model non-contact interactions such as gravity; Force Interactions, which describe how forces cause changes in motion; Change, which examines variations in quantities like velocity and energy; and Conservation, which emphasizes principles like the preservation of mechanical energy and linear momentum in isolated systems.[2] The exam assesses three interconnected science practices that foster scientific inquiry and modeling, rather than rote memorization, ensuring students demonstrate reasoning through evidence and application to novel scenarios.[2] Science Practice 1 requires students to create representations, such as free-body diagrams, motion graphs, or energy bar charts, to visualize and analyze physical systems; this practice is exclusively evaluated in 20–35% of free-response questions, where students translate between qualitative descriptions and quantitative models.[2] Science Practice 2 focuses on mathematical routines essential for physics problem-solving, including deriving symbolic expressions from physical principles (assessed in 15–20% of multiple-choice questions and 30–40% of free-response questions), performing calculations and estimations to quantify outcomes (20–25% of multiple-choice questions), and conducting comparisons or making predictions based on functional dependencies, such as how changes in one variable affect another (10–15% each in multiple-choice questions).[2] For instance, students might derive an expression for centripetal acceleration or estimate the speed of a falling object using conservation laws.[2] Science Practice 3 emphasizes scientific questioning and argumentation, where students design experiments to test hypotheses (35–45% of free-response questions), apply conceptual models to interpret data or scenarios (20–25% of multiple-choice questions), and justify scientific claims with evidence from observations or calculations (5–10% of multiple-choice questions).[2] This practice integrates experimental skills, such as outlining procedures for measuring friction coefficients or analyzing oscillatory motion, to promote evidence-based reasoning.[2] Laboratory investigations are woven into the assessment via experimental design free-response questions, which require students to propose methods, predict results, and evaluate uncertainties, reflecting the course's dedication of about 25% of instructional time to hands-on inquiry.[2] Overall, the exam prioritizes conceptual understanding and interconnected skills over isolated fact recall, aligning with college-level expectations for applying physics to real-world contexts.[2]Scoring and Performance Data
The AP Physics 1 exam is scored on a 1 to 5 scale, with scores of 3 or higher generally qualifying students for college credit or advanced placement at many institutions. The exam consists of two sections: multiple-choice questions (MCQs) and free-response questions (FRQs). Section I includes 40 MCQs, each worth 1 point for a correct answer, with no partial credit or points deducted for incorrect or unanswered questions; this section contributes 50% to the composite score. Section II comprises 4 FRQs, scored by trained readers using detailed rubrics that award points for specific elements such as correct principles, calculations, and explanations; the raw FRQ points are weighted to also contribute 50% to the composite score, which is then converted to the final 1-5 scale.[3] Historical performance data indicates that AP Physics 1 has one of the lowest pass rates (scores of 3 or higher) among all AP exams, reflecting the course's rigor in introducing algebra-based physics concepts. The mean score has hovered around 2.5 to 2.7 for most years since the exam's inception in 2015, with the 2024 mean at 2.55 and a 47.3% pass rate. Score distributions show steady improvement in pass rates over time, rising from approximately 32% in 2015 to 47.3% in 2024, attributed to increased teacher resources and student preparation. For 2025, following the addition of the fluids unit, the mean score rose to approximately 3.07, with a pass rate of 67.4%, marking a significant uptick possibly influenced by the exam's new digital format.[14][29][30]| Year | % Score 5 | % Score 4 | % Score 3 | % Score 3+ | Mean Score |
|---|---|---|---|---|---|
| 2025 | 19.8 | 24.7 | 22.9 | 67.4 | ~3.07 |
| 2024 | 10.2 | 17.9 | 19.2 | 47.3 | 2.55 |
| 2023 | 8.8 | 18.3 | 20.3 | 47.4 | 2.54 |
| 2022 | 7.9 | 17.0 | 18.5 | 43.4 | 2.47 |
| 2021 | 6.9 | 16.3 | 18.7 | 41.9 | 2.45 |