Fact-checked by Grok 2 weeks ago

Exact differential

In and physics, an exact differential is a df = P(\mathbf{x}) \cdot d\mathbf{x} in multiple variables that arises as the total of a function f, such that P_i = \frac{\partial f}{\partial x_i} for each component, ensuring the \int df between two points is path-independent and equals f(B) - f(A). For a two-variable case, df = P(x,y) \, dx + Q(x,y) \, dy is exact if there exists f(x,y) satisfying these relations. A key test for exactness, assuming the functions are continuously differentiable, is the equality of mixed partial derivatives: \frac{\partial P}{\partial y} = \frac{\partial Q}{\partial x}. This condition stems from Clairaut's theorem on the equality of mixed partials and guarantees the existence of the locally. Exact differentials play a central role in , where they distinguish state functions—like internal energy U or entropy S, whose differentials dU and dS are exact—from path functions like heat \delta Q and work \delta W, which are inexact and depend on the process path. In this context, exactness implies the function's value depends only on the system's state, not its history, enabling the formulation of fundamental relations such as dU = \delta Q - \delta W. In the study of differential equations, an equation M(x,y) \, dx + N(x,y) \, dy = 0 is termed if its left side is the exact differential of some \Psi(x,y), allowing direct to yield the implicit solution \Psi(x,y) = C. This approach simplifies solving equations without needing integrating factors, provided the exactness condition holds.

Definition and Properties

Formal Definition

In , an is a that arises as the total differential of some . Consider a differential form in two variables, \omega = P(x,y)\, dx + Q(x,y)\, dy. This form is if there exists a f(x,y) such that df = \omega, meaning \frac{\partial f}{\partial x} = P(x,y) and \frac{\partial f}{\partial y} = Q(x,y). A necessary and sufficient condition for exactness in two variables, assuming sufficient of P and Q, is that the mixed partial derivatives satisfy \frac{\partial P}{\partial y} = \frac{\partial Q}{\partial x}. This equality follows from Clairaut's theorem (also known as Schwarz's theorem), which states that if the second partial derivatives of f are continuous, then \frac{\partial^2 f}{\partial y \partial x} = \frac{\partial^2 f}{\partial x \partial y}. This concept generalizes to n variables, where a 1-form \omega = \sum_{i=1}^n P_i(\mathbf{x})\, dx_i is exact if it is the exterior derivative of a 0-form (scalar ), or equivalently, if the associated \mathbf{P} = (P_1, \dots, P_n) is conservative; in particular, in three dimensions, its vanishes: \nabla \times \mathbf{P} = \mathbf{0}. The formalization of exact differentials within the framework of occurred in the late , primarily through the independent contributions of and , who developed the modern notation and theorems for vector fields and their differentials.

Path Independence

A defining characteristic of an exact differential \omega is that the \int_C \omega between two fixed endpoints a and b in the domain is independent of the path C connecting them. This property holds because an exact differential satisfies \omega = df for some function f, ensuring the integral captures only the net change in f. The fundamental for line integrals formalizes this path independence. For a smooth C parameterized by \vec{r}(t) with a \leq t \leq b, and \omega = \nabla f \cdot d\vec{r} where \nabla f is continuous, the theorem states: \int_C \nabla f \cdot d\vec{r} = f(\vec{r}(b)) - f(\vec{r}(a)). This equality depends solely on the endpoints, not the specific path. To derive this, parameterize the curve and apply the chain rule: \nabla f \cdot \frac{d\vec{r}}{dt} = \frac{d}{dt} [f(\vec{r}(t))]. Integrating both sides from t = a to t = b and invoking the yields: \int_a^b \frac{d}{dt} [f(\vec{r}(t))] \, dt = f(\vec{r}(b)) - f(\vec{r}(a)). Thus, the line integral simplifies to the difference in the values, confirming path independence. In contrast, inexact differentials lead to path-dependent integrals, where the value varies with the chosen route. For instance, the work done by non-conservative forces, such as , depends on the taken, as dissipation accumulates differently along varied paths. A simple example illustrates this for an exact differential. Consider \omega = 2xy \, dx + x^2 \, dy, which is with potential f(x,y) = x^2 y. The from (0,0) to (1,1) equals f(1,1) - f(0,0) = 1 - 0 = 1, regardless of path. Direct computation along the straight line y = x (where dy = dx, $0 \leq x \leq 1) gives: \int_0^1 (2x \cdot x + x^2 \cdot 1) \, dx = \int_0^1 (3x^2) \, dx = 1. Along the path (first along the x-axis to (1,0), then up the y-axis to (1,1)), the is $0 + \int_0^1 x^2 \, dy = 1 (with x=1), confirming the same result.

Thermodynamic Applications

State Functions

In thermodynamics, state functions are properties of a system that depend solely on its current state, defined by variables such as , , and volume, rather than the history or path taken to reach that state. Examples include U, H, S, F, and G. In contrast, process functions like Q and work W are path-dependent, meaning their values vary with the specific process connecting initial and final states. The of a is , ensuring that changes in the are independent of the . For instance, of states that dU = \delta Q - \delta W, where \delta Q and \delta W are inexact s for and work transfers (with \delta W denoting work done by the system), but their difference dU is . This holds for any , reversible or irreversible, implying that the finite change \Delta U = U_{\text{final}} - U_{\text{initial}} depends only on the initial and final states, not the intermediate . A key criterion for identifying state functions is that the line integral of their differential around any closed cycle vanishes: \oint d\phi = 0, where \phi is the state function. This property confirms path independence and distinguishes state functions from process functions, whose cyclic integrals are generally nonzero. In modern contexts, such as non-equilibrium thermodynamics, the exactness of differentials for state functions may not hold in the traditional sense, as systems deviate from equilibrium states during irreversible processes like rapid expansions or transport phenomena. Here, internal energy and other state variables can exhibit path-dependent behaviors due to entropy production and spatial gradients, requiring extensions like local equilibrium approximations or additional flux variables to describe dynamics accurately.

Examples in Thermodynamics

In thermodynamics, the internal energy U serves as a fundamental example of an exact differential, expressed as dU = T \, dS - P \, dV, where T is , S is , P is , and V is . This form is exact because U is a depending solely on the state variables S and V, ensuring that changes in U are path-independent and determined only by the initial and final states of the system. Similarly, the H, defined as H = U + PV, has the exact differential dH = T \, dS + V \, dP. As a with natural variables S and P, H exhibits path independence, meaning the change \Delta H for any process depends only on the endpoints, making it particularly useful for constant-pressure processes where \Delta H equals the transferred. The path independence of exact differentials like those for U and H highlights a key distinction in thermodynamic processes: reversible versus irreversible. In reversible processes, the equality dU = T \, dS - P \, dV holds exactly, allowing full recovery of work and heat along the path, whereas irreversible processes involve inequalities (e.g., dq < T \, dS), but the net change in state functions such as \Delta U or \Delta H remains path-independent, relying only on initial and final states. The Gibbs free energy G = H - TS provides another illustration, with its exact differential dG = -S \, dT + V \, dP, exact due to its dependence on state variables T and P. This form is crucial in phase equilibria, where at constant T and P, the minimum G determines stable phases, and equilibrium between phases occurs when their chemical potentials are equal, ensuring dG = 0.

Mathematical Formulations

One-Dimensional Case

In the one-dimensional case, the f(x) is given by df = f'(x)\, dx, where f'(x) is the of f with respect to x. This form is exact by definition, as it directly represents the change in f corresponding to an change dx in the independent variable x. The integral of this exact differential from a point a to b yields \int_a^b df = f(b) - f(a), which follows from the and holds regardless of the specific path taken, since in one dimension there is only a single possible path along the line. This path independence is a trivial consequence of the linear nature of . A representative example from illustrates this concept: the s(t) of a particle moving along a straight line serves as the whose exact differential is the ds = v(t)\, dt, where v(t) is the . Integrating this differential gives the change in \Delta s = \int_{t_1}^{t_2} v(t)\, dt = s(t_2) - s(t_1), directly linking the of to . This one-dimensional formulation is fundamentally tied to the concept of s in , where the exact differential df = f'(x)\, dx implies that f(x) is the antiderivative of f'(x), up to a constant, ensuring that recovers the original function precisely.

Multidimensional Case

In the multidimensional case, the concept of an exact extends beyond one variable to differential forms on \mathbb{R}^n, where a 1-form \omega = P \, dx_1 + Q \, dx_2 + \cdots + R \, dx_n is exact if there exists a function f such that \omega = df, meaning the coefficients are the partial derivatives of f. This corresponds to the (P, Q, \dots, R) being conservative, with \nabla f = (P, Q, \dots, R). In two dimensions, consider the 1-form \omega = P(x,y) \, dx + Q(x,y) \, dy. The form is exact if and only if it is closed, satisfying the condition \frac{\partial P}{\partial y} = \frac{\partial Q}{\partial x}, which is equivalent to the curl of the associated vector field (P, Q) being zero: \nabla \times (P, Q) = 0. This test ensures path independence of the line integral \int_C \omega, a hallmark of conservative fields where the integral depends only on the endpoints. For three dimensions, the 1-form \omega = P(x,y,z) \, dx + Q(x,y,z) \, dy + R(x,y,z) \, dz is exact if the (P, Q, R) has zero : \nabla \times (P, Q, R) = 0. The components of the are \left( \frac{\partial R}{\partial y} - \frac{\partial Q}{\partial z}, \frac{\partial P}{\partial z} - \frac{\partial R}{\partial x}, \frac{\partial Q}{\partial x} - \frac{\partial P}{\partial y} \right) = (0, 0, 0). Again, this implies the \int_C \omega is path-independent, equaling f(B) - f(A) for endpoints A and B. To find the potential function f for an exact form, integrate one coefficient while treating others as constants, then determine remaining arbitrary functions using the other coefficients. In two dimensions, integrate P with respect to x to obtain f(x,y) = \int P \, dx + h(y), differentiate with respect to y, and set equal to Q to solve for h'(y). In three dimensions, similarly integrate P with respect to x to get f(x,y,z) = \int P \, dx + g(y,z), then use Q to find \frac{\partial g}{\partial y} and R to find \frac{\partial g}{\partial z}. Verify by checking \nabla f = (P, Q, R). A classic example is the gravitational field near Earth's surface, approximated as \mathbf{F} = -g \hat{z}, which is conservative since \nabla \times \mathbf{F} = 0, with potential f = g z + C and path-independent work done by the field. In simply connected domains, such as \mathbb{R}^2 or \mathbb{R}^3 minus isolated points, the Poincaré lemma guarantees that every closed 1-form is exact, ensuring the existence of a potential function under these topological conditions.

Differential Relations

Reciprocity Relation

In the context of , the reciprocity relation arises as a direct consequence of the exactness condition for a . Consider a Z(x, y) whose total differential is dZ = M(x, y)\, dx + N(x, y)\, dy, where the differential is exact. This exactness requires that the mixed second partial derivatives of Z are equal, specifically \frac{\partial^2 Z}{\partial x \partial y} = \frac{\partial^2 Z}{\partial y \partial x}. Since \frac{\partial M}{\partial y} = \frac{\partial^2 Z}{\partial y \partial x} and \frac{\partial N}{\partial x} = \frac{\partial^2 Z}{\partial x \partial y}, the reciprocity relation follows: \frac{\partial M}{\partial y} = \frac{\partial N}{\partial x}. This relation can be proven using the equality of mixed partial derivatives, a fundamental in stating that if the second partial derivatives are continuous, then \frac{\partial^2 Z}{\partial x \partial y} = \frac{\partial^2 Z}{\partial y \partial x}. For the exact differential dZ, differentiating M with respect to y yields the left-hand mixed partial, while differentiating N with respect to x yields the right-hand mixed partial; their equality enforces the reciprocity condition. This proof underscores the path independence of exact differentials, as the condition ensures Z is a . In , the reciprocity relation applies to state functions like U(S, V), whose differential is dU = T\, dS - P\, dV, an exact differential where M = T and N = -P. Applying the relation gives \left( \frac{\partial T}{\partial V} \right)_S = -\left( \frac{\partial P}{\partial S} \right)_V. This thermodynamic reciprocity links measurable quantities such as temperature and pressure to changes. The reciprocity relation serves as the foundation for deriving , which are additional equalities among thermodynamic partial derivatives obtained by applying the reciprocity condition to various thermodynamic potentials. For instance, from the differential of the dG = -S\, dT + V\, dP, reciprocity yields \left( \frac{\partial S}{\partial P} \right)_T = -\left( \frac{\partial V}{\partial T} \right)_P, facilitating experimental determination of thermodynamic properties.

Cyclic Relation

In multivariable calculus, the cyclic relation, also known as the triple product rule or Euler's chain rule, emerges as a consequence of the exactness of a total differential for interdependent variables x, y, and z, where z = z(x, y). The relation states that \left( \frac{\partial x}{\partial y} \right)_z \left( \frac{\partial y}{\partial z} \right)_x \left( \frac{\partial z}{\partial x} \right)_y = -1. This identity holds because the partial derivatives must satisfy consistency conditions for the differentials to be path-independent. The derivation follows directly from the total differential dz = \left( \frac{\partial z}{\partial x} \right)_y dx + \left( \frac{\partial z}{\partial y} \right)_x dy, which is if the mixed partial derivatives are equal, i.e., \frac{\partial^2 z}{\partial y \partial x} = \frac{\partial^2 z}{\partial x \partial y}. To obtain the cyclic form, express y as a of x and z, yielding dy = \left( \frac{\partial y}{\partial x} \right)_z dx + \left( \frac{\partial y}{\partial z} \right)_x dz. Substituting the expression for dz and collecting terms leads to coefficients that must vanish for arbitrary dx and dy, resulting in the product equaling -1. This ensures the is integrable and the variables are related through a . In thermodynamics, the cyclic relation applies to state functions derived from exact differentials like the internal energy differential dU = T \, dS - P \, dV, where U = U(S, V), T = \left( \frac{\partial U}{\partial S} \right)_V, and P = -\left( \frac{\partial U}{\partial V} \right)_S. For instance, considering the interdependent variables P, V, and T, the relation takes the form \left( \frac{\partial P}{\partial T} \right)_V \left( \frac{\partial T}{\partial V} \right)_P \left( \frac{\partial V}{\partial P} \right)_T = -1, which links the coefficients T and P to measurable properties like the thermal expansion coefficient \alpha = \frac{1}{V} \left( \frac{\partial V}{\partial T} \right)_P and isothermal compressibility \beta = -\frac{1}{V} \left( \frac{\partial V}{\partial P} \right)_T, yielding \left( \frac{\partial P}{\partial T} \right)_V = \frac{\alpha}{\beta}. This form is verified for ideal gases using the equation of state PV = nRT. A common misconception is applying the cyclic relation to non-state variables, such as heat Q or work W, whose differentials dQ and dW are inexact and path-dependent; the relation holds only for state functions where the total differential is , ensuring the partial derivatives reflect intrinsic dependencies rather than process-specific paths. The cyclic relation is distinct from but complementary to the reciprocity relation, which involves pairwise in second derivatives.

Derived Equations and Identities

In Two Dimensions

In two dimensions, an exact differential takes the form \omega = M(x,y) \, dx + N(x,y) \, dy, where there exists a function f(x,y) such that \omega = df. The total differential of f is given by df = \frac{\partial f}{\partial x} \, dx + \frac{\partial f}{\partial y} \, dy, so M = \partial f / \partial x and N = \partial f / \partial y. Differentiating M yields dM = \frac{\partial M}{\partial x} \, dx + \frac{\partial M}{\partial y} \, dy, and since M = \partial f / \partial x, the equality of mixed partial derivatives implies \partial M / \partial y = \partial^2 f / \partial y \partial x = \partial N / \partial x. This leads to the key integrability condition for exactness: \partial M / \partial y = \partial N / \partial x. If the differential is inexact, an \mu(x,y) may exist such that \mu \omega becomes exact, satisfying \partial (\mu M) / \partial y = \partial (\mu N) / \partial x, though the focus here remains on exact cases. A prominent application arises in , where the enthalpy differential dH = T \, [dS](/page/DS) + V \, [dP](/page/DP) is exact, with T = (\partial H / \partial S)_P and V = (\partial H / \partial P)_S. The exactness condition \partial T / \partial P = \partial V / \partial S then yields the Maxwell relation \left( \frac{\partial T}{\partial P} \right)_S = \left( \frac{\partial V}{\partial S} \right)_P. The reciprocity relation in two dimensions, stemming from the equality of mixed partials, ensures that the Jacobian matrix of the transformation defined by the partial derivatives is symmetric. For df = M \, dx + N \, dy, the Jacobian J = \begin{pmatrix} \frac{\partial M}{\partial x} & \frac{\partial M}{\partial y} \\ \frac{\partial N}{\partial x} & \frac{\partial N}{\partial y} \end{pmatrix} satisfies J^T = J due to \partial M / \partial y = \partial N / \partial x.

Generalizations to Higher Dimensions

In the context of and , the concept of an exact differential extends naturally to n dimensions through the framework of differential forms. A 1-form \omega on an open subset of \mathbb{R}^n is said to be exact if there exists a f: \mathbb{R}^n \to \mathbb{R} such that \omega = df; equivalently, \omega is closed if its vanishes, i.e., d\omega = 0. On contractible domains, such as star-shaped open sets in \mathbb{R}^n, the guarantees that every closed form is exact, ensuring the existence of a whose yields the form. This generalizes the two-dimensional condition where exactness implies path-independent line integrals, now holding for higher-dimensional integrals over simply connected regions. From such exact differentials arise higher-order relations analogous to Maxwell's relations in thermodynamics, derived from the equality of mixed partial derivatives of the potential function. For a thermodynamic potential \Phi depending on n independent variables x_1, \dots, x_n, the exactness of d\Phi = \sum_{i=1}^n y_i \, dx_i (where the y_i are conjugate variables) implies that the Hessian matrix of second partial derivatives is symmetric, yielding \frac{\partial^2 \Phi}{\partial x_i \partial x_j} = \frac{\partial^2 \Phi}{\partial x_j \partial x_i} for all i, j. This produces \binom{n}{2} independent Maxwell-like relations connecting cross derivatives, such as \left( \frac{\partial y_i}{\partial x_j} \right) = \left( \frac{\partial y_j}{\partial x_i} \right), enforcing the symmetry of the Hessian and reducing the number of independent second derivatives from n^2 to \frac{n(n+1)}{2}. These identities facilitate relating measurable quantities like pressure and entropy across multiple state variables. A concrete example occurs in three-dimensional thermodynamics with the F(T, V, N), where dF = -S \, dT - P \, dV + \mu \, dN and S, P, \mu are , , and , respectively. Exactness leads to cross-partial equalities, including \left( \frac{\partial S}{\partial V} \right)_{T,N} = \left( \frac{\partial P}{\partial T} \right)_{V,N}, \left( \frac{\partial S}{\partial N} \right)_{T,V} = -\left( \frac{\partial \mu}{\partial T} \right)_{V,N}, and \left( \frac{\partial P}{\partial N} \right)_{T,V} = \left( \frac{\partial \mu}{\partial V} \right)_{T,N}, enabling derivations of phase behavior and response functions in multi-component systems. The condition for closedness in n dimensions corresponds to the vanishing of the n-dimensional analogue of the curl for the associated vector field. For a 1-form \omega = \sum_{i=1}^n F_i \, dx_i, d\omega = 0 implies that the 2-form components satisfy \sum_{i<j} \left( \frac{\partial F_j}{\partial x_i} - \frac{\partial F_i}{\partial x_j} \right) dx_i \wedge dx_j = 0, or in vector terms, the antisymmetric part of the Jacobian vanishes pairwise. However, on non-contractible manifolds, closed forms need not be exact; de Rham cohomology quantifies this obstruction, with cohomology groups H^k(M) measuring the dimension of closed k-forms modulo exact ones, as seen in examples like the punctured plane where the angular form d\theta is closed but not exact.

References

  1. [1]
    Exact Differential -- from Wolfram MathWorld
    A differential of the form df=P(x,y)dx+Q(x,y)dy is exact (also called a total differential) if intdf is path-independent.
  2. [2]
    9.2: Exact and Inexact Differentials - Chemistry LibreTexts
    Jul 3, 2024 · If a differential is the total differential of a function, we will call the differential exact. What we did so far is correct, but it is not the ...
  3. [3]
    Differential Equations - Exact Equations - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will discuss identifying and solving exact differential equations. We will develop of a test that can be used to identify ...
  4. [4]
  5. [5]
    Calculus III - Higher Order Partial Derivatives
    Nov 16, 2022 · Example 2 Verify Clairaut's Theorem for f(x,y)=xe−x2y2 f ( x , y ) = x e − x 2 y 2 . We'll first need the two first order derivatives. Now, ...
  6. [6]
    6.5 Divergence and Curl - Calculus Volume 3 | OpenStax
    Mar 30, 2016 · With the next two theorems, we show that if F is a conservative vector field then its curl is zero, and if the domain of F is simply connected ...
  7. [7]
    2 Differential Calculus of Vector Fields - Feynman Lectures - Caltech
    The interesting theorem is that if the curl A is zero, then A is always the gradient of something—there is some scalar field ψ (psi) such that A is equal to gra ...
  8. [8]
    Oliver Heaviside - Biography - MacTutor - University of St Andrews
    His operational calculus, developed between 1880 and 1887, caused much controversy however. He introduced his operational calculus to enable him to solve the ...
  9. [9]
    4.5: Path Independence, Conservative Fields, and Potential Functions
    Oct 27, 2024 · Theorem 1: Fundamental Theorem of Line Integrals. Let C be a smooth ... line integral depends only on the two endpoints, not on the path.
  10. [10]
    Calculus III - Fundamental Theorem for Line Integrals
    Nov 16, 2022 · The theorem states that for a smooth curve C, ∫C∇f⋅d→r = f(→r(b))−f(→r(a)), where the path doesn't matter, and the integral is independent of  ...
  11. [11]
    2.1: Work and the Inexact Differential - Chemistry LibreTexts
    Jan 19, 2025 · For example, you can work efficiently or inefficiently ... Work is path-dependent because it can be performed efficiently or inefficiently.
  12. [12]
    8.2 Conservative and Non-Conservative Forces | University Physics ...
    The work done by a non-conservative force depends on the path taken. Equivalently, a force is conservative if the work it does around any closed path is zero:.
  13. [13]
    4.1. Differentials — Thermodynamics and Kinetics 0.1 documentation
    An exact differential yields the same integral regardless of path. It turns out that state functions have exact differentials.
  14. [14]
    Chapter 3 First Law: The Machinery
    Recall that change in internal energy, dU, is called an exact differential because it depends only on initial and final state of the system but not the path. We ...
  15. [15]
    [PDF] Entropy
    o The integral of a state variable around a closed path is zero. ○ Mathematically, this means that differentials of state functions are exact differentials.
  16. [16]
    [PDF] STATISTICAL MOLECULAR THERMODyNAMICS
    A key property of a state function is that its differential can be integrated in a normal, path independent way. Work and heat are not state functions. w = − P.
  17. [17]
    [PDF] Thermodynamics - Oregon State University
    Dec 14, 2003 · 6.3 Non-equilibrium Thermodynamics. ... these Free Energy state functions have exact differentials, and as a result, lead.
  18. [18]
    None
    ### Summary of Exact Differentials in Thermodynamics
  19. [19]
    [PDF] 553 Chapter 16. Foundations of Thermodynamics
    The internal energy is the heat transfer at constant volume, the enthalpy is the heat transfer at constant pressure, the Helmholtz energy is the maximum total ...
  20. [20]
    4. Free Energy and Equilibrium
    ... Gibbs free energy. Differentials of internal energy and enthalpy, SOLUTION. Answers: For internal energy the differential form of the first law is. dU=dq+dw.
  21. [21]
    Binary Phase Equilibria Tutorial - CalTech GPS
    At constant pressure and entropy, equilibrium is found at the minimum in enthalpy H = E + PV. For a phase, the Gibbs free energy is a function of P, T, and ...
  22. [22]
    [PDF] 2 Grad, div, curl
    a small error that goes to zero faster than |kr − kr0| as kr → kr0. The limit as kr → kr0 can be written as the exact differential relation df(kr) = dkr ...<|separator|>
  23. [23]
    [PDF] Chapter 4 One Dimensional Kinematics - MIT OpenCourseWare
    Because the velocity is the derivative of position with respect to time, the x -component of the acceleration is the second derivative of the position function,.
  24. [24]
    [PDF] Vector fields and differential forms - Arizona Math
    Sep 25, 2008 · DIFFERENTIAL 1-FORMS. 3. In two dimensions an exact differential form is of the form dh(x, y) = ∂h(x, y). ∂x dx +. ∂h(x, y). ∂y dy. (1.8). If ...
  25. [25]
    Calculus III - Conservative Vector Fields - Pauls Online Math Notes
    Nov 16, 2022 · ... finding a potential function for the vector field? The first question is easy to answer at this point if we have a two-dimensional vector field.
  26. [26]
    2.2 Exact one-forms and conservative vector fields
    We introduce a class of one-forms that arise as differentials of a function; the associated vector field is the gradient of the function.
  27. [27]
    4.6 Exact and closed k -forms
    However, there is a simple case for which closed k -forms are always exact, as in Theorem 3.6.1. This is called “Poincare's lemma” for k -forms.
  28. [28]
    None
    ### Summary of Euler Reciprocity Relation for Exact Differentials
  29. [29]
    [PDF] ESCI 341 – Atmospheric Thermodynamics
    One property of exact differentials is that if df(x,y) = Mdx +Ndy, then x. N y. M. ∂. ∂. = ∂. ∂ . Euler reciprocity relation. ○ This allows us to write, from ...
  30. [30]
    [PDF] MATH BACKGROUND FOR THERMODYNAMICS ∑ - SMU Physics
    in the various thermodynamic state functions are exact differentials, while the work done on the system dw and the heat absorbed by the system dq during the ...
  31. [31]
    [PDF] Chapter 11: Thermodynamic Property Relations - KFUPM
    In this section we develop general relations for changes in internal energy, enthalpy, and entropy in terms of pressure, specific volume, temperature, and ...
  32. [32]
    [PDF] Thermodynamic Relations
    dQ = du + pdv. According to second law, ds = dQ. T. F. HGIKJrev. Combining these ... Using the cyclic relation for p, v, T which is. ∂. ∂. F. H. I. K. ∂. ∂. F.
  33. [33]
    [PDF] Intensive and extensive properties • State function a
    Eq 15 is called the Euler chain relation (or the cyclic rule), applicable only for state functions.
  34. [34]
    [PDF] The Exact Form and General Integrating Factors
    In this chapter, we will develop a more general approach to converting a differential equation to a form (the “exact form”) that can be integrated through a ...
  35. [35]
    [PDF] 8.044s13 Maxwell Relations: A Wealth of Partial Derivatives
    If one considers E to be a function of S and V (its “natural' variables), then one can expand it as an exact differential. dE(S, V ) = ∂E. ∂S. dS +. dV. V. ∂E.
  36. [36]
    lecture 7
    This is one of Maxwell's relations. It is a result of the fact that dE is an exact differential. It states the relation that the macroscopic parameters in eq. ( ...
  37. [37]
    [PDF] Differential Forms and Integration - UCLA Mathematics
    The concept of a closed form corresponds to that of a conservative force in physics (and an exact form corresponds to the concept of having a potential function) ...
  38. [38]
  39. [39]
    [PDF] The Maxwell relations
    The Maxwell relations. A number of second derivatives of the fundamental relation have clear physical significance and can be measured experimentally. For ...Missing: exact | Show results with:exact