Fact-checked by Grok 2 weeks ago

Initial condition

An initial condition refers to a set of values that specify the state of a —such as the , , or other variables—at a designated starting point, typically time t = 0, in the context of s describing dynamical evolution. These conditions are essential for transforming a general , which yields a family of solutions, into a unique particular solution that accurately models the 's behavior from the outset. In mathematical terms, for an (ODE) of order n, exactly n initial conditions are required, often involving the and its up to order n-1 at the initial point. In equations, conditions form the core of an (IVP), where the goal is to solve \frac{dy}{dt} = f(t, y) subject to y(t_0) = y_0, ensuring the solution passes through the specified point. For higher-order equations, such as second-order modeling physical motion, conditions might include both and , as in y(0) = y_0 and y'(0) = v_0. This framework underpins applications in physics, including under gravity, where the IVP v' = -9.8 with v(0) = 20 yields the v(t) = 20 - 9.8t in meters per second. For partial differential equations (PDEs), which involve multiple variables like space and time, initial conditions specify the system's state across the spatial domain at the initial time, distinct from boundary conditions that constrain behavior on spatial edges. In , initial conditions determine the trajectory in according to \dot{\vec{x}} = \vec{f}(\vec{x}), with uniqueness guaranteed under of \vec{f}. A critical aspect arises in chaotic systems, where even infinitesimal perturbations in initial conditions lead to exponentially diverging outcomes, quantified by positive Lyapunov exponents \lambda > 0 such that |\delta(t)| \approx \delta_0 e^{\lambda t}, limiting long-term predictability despite deterministic rules.

Fundamentals

Definition

In , physics, and , an initial condition refers to the specified state or values of variables in a at the start of a time interval, which uniquely determine the system's subsequent evolution according to its governing equations. This concept is fundamental to initial value problems (IVPs), where the initial condition ensures the existence and uniqueness of solutions under suitable assumptions, such as those provided by the Picard-Lindelöf theorem for ordinary differential equations (ODEs). For a single first-order ODE of the form \frac{dy}{dt} = f(t, y) with y(t_0) = y_0, the initial condition y(t_0) = y_0 fixes the solution's value at the initial time t_0, allowing the trajectory to be traced forward or backward in time. In higher-order ODEs or systems of n first-order equations, multiple initial conditions—corresponding to the solution and its derivatives up to order n-1, or equivalently n state variables—are required to specify a unique solution. For instance, in the second-order equation \frac{d^2y}{dt^2} + p(t)\frac{dy}{dt} + q(t)y = g(t), initial conditions might be y(t_0) = y_0 and y'(t_0) = y_0'. In the broader context of dynamical systems, initial conditions represent the starting point in , influencing whether the system follows a periodic , converges to an , or exhibits behavior depending on the system's nonlinearity and parameters. Discrete-time systems, such as those governed by difference equations like x_{k+1} = f(x_k), similarly require an value x_0 to generate the sequence \{x_k\}. These conditions are not merely mathematical artifacts but encode empirical data, such as and in Newtonian , to model real-world phenomena accurately.

Role in Dynamical Systems

In , initial conditions specify the state of the system at a reference time, typically t = 0, and play a pivotal role in determining the unique or followed thereafter. For a system governed by an (ODE) of the form \dot{x} = f(x), where x \in \mathbb{R}^n and f is a , the (IVP) is completed by setting x(0) = x_0, which defines the starting point in the . This specification ensures that the solution x(t) evolves as the \phi_t(x_0), tracing the system's path forward and backward in time under appropriate conditions. The fundamental importance of initial conditions lies in theorems guaranteeing existence and uniqueness of solutions. The Picard-Lindelöf theorem states that if f is locally continuous in x, then for any initial condition x_0, there exists a unique local to the IVP on some interval around t = 0. This condition, often satisfied by C^1 vector fields, prevents non-uniqueness issues, such as those arising in non- cases like \dot{x} = |x|^{1/2} with x(0) = 0, where multiple emanate from the same initial point. Global existence follows if the remains bounded, ensuring the trajectory is defined for all time. These properties underpin the predictability of system evolution from given initial states. Beyond , initial conditions influence the qualitative behavior of trajectories, particularly in nonlinear systems. In or dynamical systems, solutions exhibit continuous dependence on initial conditions: small perturbations in x_0 yield correspondingly small deviations in \phi_t(x_0) over finite times, as quantified by the shadowing lemma in Anosov diffeomorphisms. However, in chaotic systems, sensitive dependence on initial conditions prevails, where infinitesimal changes in x_0 lead to exponentially diverging trajectories, measured by positive Lyapunov exponents \chi(x, v) = \lim_{t \to \infty} \frac{1}{t} \log \|\mathrm{D}\phi_t(x) v\| > 0. The , \dot{x} = \sigma(y - x), \dot{y} = x(\rho - z) - y, \dot{z} = xy - \beta z, exemplifies this, as nearby initial conditions rapidly separate, limiting long-term predictability despite short-term accuracy. This sensitivity underscores the role of initial conditions in distinguishing deterministic from , a central to applications in and . For discrete-time systems, such as iterations of a x_{n+1} = f(x_n), initial conditions similarly dictate the \{f^n(x_0)\}, with expansive maps like the horseshoe demonstrating akin to continuous cases. Overall, initial conditions bridge the system's equations to its observable dynamics, enabling analysis of attractors, , and bifurcations.

Linear Systems

Discrete-Time Systems

In discrete-time linear systems, the dynamics are typically represented in state-space form as x(k+1) = A x(k) + B u(k), where x(k) \in \mathbb{R}^n is the at time step k, u(k) \in \mathbb{R}^m is the input vector, A \in \mathbb{R}^{n \times n} is the , and B \in \mathbb{R}^{n \times m} is the input matrix. The output is given by y(k) = C x(k) + D u(k), with C \in \mathbb{R}^{p \times n} and D \in \mathbb{R}^{p \times m}. The initial condition x(0) = x_0 encapsulates the system's starting state and determines the natural response, influencing the entire trajectory even in the absence of inputs. The general solution for the state trajectory decomposes into a homogeneous part driven by the initial condition and a particular part due to inputs: x(k) = A^k x_0 + \sum_{i=0}^{k-1} A^{k-1-i} B u(i). If A is diagonalizable as A = T \Lambda T^{-1} with \Lambda containing eigenvalues \lambda_i, then A^k = T \Lambda^k T^{-1}, where \Lambda^k = \operatorname{diag}(\lambda_1^k, \dots, \lambda_n^k), allowing explicit computation of the initial condition's contribution. This formulation highlights how x_0 propagates through powers of A, with the system's behavior depending on the spectral properties of A. For zero input u(k) = 0, the response simplifies to x(k) = A^k x_0, underscoring the initial condition's role in unforced evolution. Initial conditions play a critical role in stability analysis. A discrete-time linear system is asymptotically stable if all eigenvalues of A have magnitudes strictly less than 1 (i.e., inside the unit in the z-plane), ensuring x(k) \to 0 as k \to \infty for any finite x_0 and zero input. In this case, the effect of the initial condition decays exponentially, with the decay rate governed by the eigenvalue with the largest . If any eigenvalue has exactly 1 (on the unit , non-repeated), the system is marginally , and the state remains bounded but may not converge to zero, depending on the of x_0 onto the corresponding eigenspace; repeated poles on the unit lead to instability due to growth. For bounded-input bounded-output ( with nonzero initial conditions, the system requires all poles of the to lie inside the unit , but internal stability (asymptotic) is assessed via the unforced response from x_0. In and , initial conditions interact with system properties to determine achievable states or reconstructible information. The system is controllable if, for any x_0 and desired x_f, there exists an input sequence steering the state from x_0 to x_f in finite steps, verified by the condition \operatorname{rank}([B, AB, \dots, A^{n-1}B]) = n. allows reconstruction of x_0 from output measurements, with the observability \operatorname{rank}([C; CA; \dots; CA^{n-1}]) = n. Nonzero initial conditions can amplify transient effects in simulations or digital control, necessitating careful specification to avoid numerical issues in implementations like sampled-data systems, where models approximate continuous dynamics via A = e^{A_c h} and initial states aligned at sampling instants.

Continuous-Time Systems

In continuous-time linear systems, the dynamics are typically modeled using state-space representations of the form \dot{x}(t) = A x(t) + B u(t), where x(t) \in \mathbb{R}^n is the state vector, u(t) \in \mathbb{R}^m is the input, and A \in \mathbb{R}^{n \times n}, B \in \mathbb{R}^{n \times m} are constant matrices for linear time-invariant (LTI) systems. The output is given by y(t) = C x(t) + D u(t), with C \in \mathbb{R}^{p \times n} and D \in \mathbb{R}^{p \times m}. The initial condition x(0) = x_0 specifies the starting state at time t = 0, which is essential for determining the system's evolution under the given input. The solution to this initial value problem is unique and given by the variation of constants formula: x(t) = e^{A t} x_0 + \int_0^t e^{A (t - \tau)} B u(\tau) \, d\tau, where e^{A t} is the state transition matrix, satisfying \frac{d}{dt} e^{A t} = A e^{A t} with e^{A \cdot 0} = I. This matrix can be computed via the matrix exponential, such as through Taylor series expansion e^{A t} = \sum_{k=0}^\infty \frac{(A t)^k}{k!} or diagonalization if A is diagonalizable. The first term, e^{A t} x_0, represents the zero-input (homogeneous) response driven solely by the initial condition, while the integral term captures the zero-state (forced) response due to the input. Uniqueness follows from the Picard-Lindelöf theorem, which guarantees a unique solution for Lipschitz continuous right-hand sides, a property satisfied by linear systems. The initial condition plays a pivotal role in the system's transient behavior and overall response. For instance, in the absence of input (u(t) = 0), the state evolves purely as x(t) = e^{A t} x_0, determining the natural modes of the system based on the eigenvalues of A. If A is Hurwitz (all eigenvalues have negative real parts), the response converges to zero regardless of x_0, but the initial condition influences the speed and path of this decay. In frequency-domain analysis via Laplace transforms, the initial condition contributes to the state as X(s) = (sI - A)^{-1} x_0 + (sI - A)^{-1} B U(s), highlighting its additive effect on the transform of the input. This separation underscores how initial conditions encode the system's "memory," enabling predictions of stability, controllability, and observability. For linear time-varying systems, where A = A(t) and B = B(t), the solution generalizes to x(t) = \Phi(t, 0) x_0 + \int_0^t \Phi(t, \tau) B(\tau) u(\tau) \, d\tau, with the state transition matrix \Phi(t, \tau) satisfying \frac{\partial}{\partial t} \Phi(t, \tau) = A(t) \Phi(t, \tau) and \Phi(\tau, \tau) = I. Here, the initial condition similarly drives the homogeneous part, but the time-varying nature complicates explicit computation, often requiring numerical methods. Stability and still hold under mild conditions on A(t), such as continuity.

Nonlinear Systems

Attractors and Basins of Attraction

In nonlinear dynamical systems, an is a compact set in the to which a significant portion of trajectories converge as time evolves to . This occurs for initial conditions within a specific , highlighting the of initial states in determining long-term behavior. The concept was formalized in the context of strange attractors by Ruelle and Takens in their 1971 study on turbulence, where they described sets exhibiting fractal geometry and chaotic dynamics that attract nearby orbits despite their complexity. The of attraction for a given comprises the set of all initial conditions whose forward orbits approach that asymptotically. In nonlinear systems, multiple attractors can coexist, each with its own , leading to multistability where the system's fate depends sensitively on the precise initial condition. For instance, in systems with symmetric potentials, such as a double-well oscillator, initial conditions near one minimum may lead to convergence to that stable equilibrium, while those on the opposite side attract to the other, with the basin boundary often forming a . A hallmark of nonlinear systems is the potential for basin boundaries, where small perturbations in initial conditions near the can cause trajectories to diverge to different attractors, amplifying in predictions. This structure arises from mechanisms, such as stretching and folding in the , and is quantified by metrics like the uncertainty exponent, which measures the scaling of misprediction probability with resolution; typical values around 0.2 indicate high sensitivity. In the forced damped , for example, the basin exhibits a box-counting of approximately 1.8, illustrating how initial conditions straddling this irregular frontier lead to unpredictable outcomes between periodic and attractors. The provides a seminal example of a chaotic and its . Defined by the equations \begin{align*} \dot{x} &= \sigma (y - x), \\ \dot{y} &= x (\rho - z) - y, \\ \dot{z} &= xy - \beta z, \end{align*} with classical parameters \sigma = 10, \rho = 28, \beta = 8/3, the is a butterfly-shaped strange with a Lyapunov around 2.06. Its encompasses nearly all initial conditions except those on the z-axis, the of the unstable origin fixed point, where trajectories approach the origin rather than the chaotic . Thus, typical initial states in the , perturbed slightly off the z-axis, converge to this single chaotic , underscoring how initial conditions outside trivial sets dictate the emergence of complex, non-periodic motion. In systems with riddled basins, such as those involving symmetry-breaking bifurcations, the basins of different attractors interpenetrate densely, with points from one basin arbitrarily close to another, further emphasizing the fragility of initial condition selection in nonlinear dynamics. This phenomenon, explored in coupled oscillator networks, implies that noise or measurement errors can readily shift trajectories across basins, impacting applications from climate modeling to neural networks.

Sensitivity and Chaos

In nonlinear dynamical systems, to initial conditions refers to the phenomenon where infinitesimally close starting points in evolve into trajectories that diverge exponentially over time, a hallmark of . This property implies that even minuscule perturbations in the initial state can lead to vastly different long-term outcomes, rendering precise long-term predictions impossible despite the system's deterministic nature. Edward Lorenz first demonstrated this in 1963 while modeling atmospheric using a simplified set of nonlinear equations, now known as the Lorenz equations, where he observed that rounding values from six to three decimal places caused trajectories to separate rapidly after an initial agreement period. The concept gained widespread recognition through the "butterfly effect," a term popularized by Lorenz to illustrate how a small change, such as the flap of a butterfly's wings in , could theoretically influence the formation of a in by amplifying through the system's dynamics. In his presentation, Lorenz emphasized that this sensitivity arises not from randomness but from the inherent structure of certain nonlinear systems, where initial differences grow at rates that double the separation distance roughly every unit of time in highly regimes. This effect underscores the practical limits of in fields like , as computational approximations inevitably introduce small errors that amplify unpredictably. Sensitivity to initial conditions is quantitatively characterized by Lyapunov exponents, which measure the average exponential rates of divergence or convergence of nearby trajectories along different directions in . Originating from the multiplicative ergodic , these exponents exist for almost all initial conditions in ergodic systems and indicate when the largest exponent is positive, signifying overall expansion in the . For instance, in the Lorenz , the maximal is approximately 0.906, confirming exponential with a characteristic of about 0.77 units, while the sum of all exponents equals the negative of the Jacobian, ensuring volume contraction onto the . This framework, applied across systems like the or , highlights how coexists with bounded, attractors despite the profound sensitivity.

Empirical Contexts

In Scientific Modeling

In scientific modeling, initial conditions refer to the specified values of state variables at the start of a , which are essential for defining the of a system's under the model's governing equations. These conditions are particularly vital in numerical simulations of physical, biological, and environmental processes, where they influence the accuracy, stability, and computational efficiency of predictions. For instance, in or climate simulations, improper initial conditions can lead to numerical instability or divergent results, necessitating careful selection based on observational data or equilibrium assumptions. The significance of initial conditions lies in their role in capturing the system's starting state, which determines subsequent behavior in both deterministic and models. In simulations of complex systems, even minor variations in initial conditions can amplify over time due to nonlinear interactions, highlighting the need for analyses to assess robustness. A study on models emphasizes that initial conditions channel interactions between components toward specific outcomes, as seen in cellular automata where differing starting configurations led to occupation rates varying from 88 to 191 squares in a tissue-growth model. Similarly, in hydrologic modeling, initial water table depths (e.g., 1 m vs. 7 m) caused surface runoff variations of 40–88% and surface storage changes of 30–78%, with dry conditions exhibiting greater persistence and requiring longer spin-up periods (up to 10 years for ). In , initial conditions affect parameter in dynamic models, where problematic values (e.g., zeros) can render parameters unobservable, impacting model validation. A systematic of models like JAK/STAT and Epo signaling pathways revealed that adjusting initial conditions via sensitivity matrix rank analysis (using ) and symbolic methods (e.g., derivatives) improved , identifying up to five unidentifiable parameters in real-world cases. For reduced-order models of high-dimensional dynamical , deriving initial conditions through normal form transformations ensures faithful reproduction of long-term behavior, as demonstrated in examples where this approach aligned low-dimensional simulations with original . Climate modeling underscores the dual influence of initial conditions and boundary forcings on predictability, with initial-condition ensembles revealing their dominance in short- to medium-range forecasts. In IPCC evaluations, climate models often initialize atmospheric states from reanalysis , but subsurface components like temperatures require spin-up to mitigate drift, affecting global and regional projections. Large initial-condition ensembles (e.g., 40+ members) have shown value in quantifying uncertainty, though they remain computationally intensive and prone to biases if not paired with observational constraints. Overall, best practices in scientific modeling involve testing multiple initial configurations, incorporating for , and evaluating spin-up to balance fidelity and efficiency across disciplines.

Philosophical Implications

The concept of initial conditions in dynamical systems raises profound questions about and predictability in the . Classical , as articulated by , posits that a —often termed —could predict the entire future state of the universe if it knew the precise initial conditions of all particles and the governing laws of nature. However, this view assumes perfect knowledge and computation, which challenges through sensitive dependence on initial conditions (SDIC), where arbitrarily small differences in starting states lead to exponentially diverging trajectories in nonlinear systems. Despite maintaining ontological —outcomes remain law-governed—SDIC renders long-term point predictions epistemically impossible, even with near-perfect initial data, as uncertainties amplify rapidly. Chaos theory further implies a novel form of unpredictability: approximate probabilistic irrelevance of the past. In chaotic systems defined by mixing properties, knowledge of sufficiently distant initial conditions becomes irrelevant for forecasting future events, as the system's evolution erases informational traces of the starting state, yielding uniform probability distributions over outcomes. This undermines Laplacean predictability without invoking indeterminism, highlighting an epistemological gap between deterministic laws and practical foresight; for instance, in the tent map model, initial distributions spread to cover the entire state space, making prior states probabilistically uninformative after a few iterations. Philosophers argue this reveals that determinism does not entail predictability, shifting focus from absolute foreknowledge to probabilistic or ensemble-based modeling in complex systems like weather or biological populations. In and physical , conditions complicate the distinction between necessary laws and contingent facts. Traditional views treat laws as modally robust while conditions are freely assignable, but in universes with a singular origin like the , such conditions may be constrained by nomic necessities to ensure consistency, such as avoiding closed timelike curves in . This raises implications for explanatory asymmetry, as seen in the second law of , where low-entropy conditions appear fine-tuned yet contingent, prompting debates on whether they reflect deeper necessities or mere happenstance. Overall, conditions thus bridge and , challenging reductionist and emphasizing the role of in scientific .

References

  1. [1]
    Initial-Value Problems | Calculus II - Lumen Learning
    An initial-value problem will consists of two parts: the differential equation and the initial condition. The differential equation has a family of solutions, ...
  2. [2]
  3. [3]
    Initial Value Problem (Initial Condition) - Statistics How To
    An initial condition is a starting point; Specifically, it gives dependent variable values (or one of its derivatives) for a certain independent variable.
  4. [4]
    [PDF] Chapter 7 Chaos and Non-Linear Dynamics - MIT OpenCourseWare
    By “sensitive to initial conditions” we mean that trajectories that start nearby initially, separate exponentially fast. Defining δ(t) as the difference between ...<|control11|><|separator|>
  5. [5]
    Differential Equations - Definitions - Pauls Online Math Notes
    Nov 16, 2022 · So, in other words, initial conditions are values of the solution and/or its derivative(s) at specific points. As we will see eventually, ...
  6. [6]
    [PDF] Introduction Differential Equations
    Definition An initial value problem is an a differential equation with a spec- ified value of the solution provided. Such a value is called an initial condition ...
  7. [7]
    [PDF] Introduction to Differential Equations, Initial Value Problems, and ...
    An initial condition is a constraint of the form y(x0) = y0, which specifies a point (x0,y0) that the unknown function must pass through.
  8. [8]
    [PDF] Chapter 4 Differential Equations
    We call this specification an initial condition. The differential equations together with an initial condition is called an initial value problem. Each initial.
  9. [9]
    Initial dynamical systems exploration - Math Insight
    We call this initial value the initial condition of the dynamical system. We often write the initial condition as a(0)=a0,. so we can use a(0) and a0 ...
  10. [10]
    [PDF] MODELING WITH DIFFERENTIAL EQUATIONS
    Definition. An Initial Condition for a differential equation is a statement of given values for the dependent and independent variables of the solution ...
  11. [11]
    [PDF] Introduction to Dynamical Systems John K. Hunter - UC Davis Math
    Solutions of chaotic systems are sensitive to small changes in the initial conditions, and Lorenz used this model to discuss the unpredictability of weather ...
  12. [12]
    [PDF] Introduction to Dynamical Systems - Ceremade
    ... dynamical systems often appear to be chaotic in that they have sensitive dependence on initial conditions, i.e., minor changes in the initial state lead to.
  13. [13]
    [PDF] Dynamical Systems Theory - University of California, Santa Barbara
    1.2 Nonlinear Dynamical Systems Theory . . . . . . . . . . . . . . . 11 ... With these initial conditions specified, the initial value problem (IVP) (1.1).
  14. [14]
    [PDF] Automatic Control 1 - Discrete-time linear systems
    Discrete-time linear system... x(k + 1) = Ax(k) + Bu(k) y(k) = Cx(k) + Du(k) x(0) = x0. Given the initial condition x(0) and the input sequence u(k), k ...
  15. [15]
    [PDF] Stability of Discrete-time systems - F.L. Lewis
    Oct 7, 2008 · Discrete-time systems are stable if poles are inside the unit circle. Asymptotically stable if poles are strictly inside, and marginally stable ...
  16. [16]
    [PDF] Chapter Five - Linear Systems
    solution with initial condition x(0) = αx01 + βx02 is given by x(t) = eAt ... (b) Show that a discrete-time linear system is asymptotically stable if and only if.
  17. [17]
    [PDF] Continuous-time linear state-space models - MIT OpenCourseWare
    In summary, knowing n solutions of the homogeneous system for n independent initial conditions, we are able to construct the general solution of this linear ...<|control11|><|separator|>
  18. [18]
    [PDF] Lectures on Linear Systems Theory - University of Notre Dame
    Dec 2, 2024 · equation for all t ≥ 0 and that satisfy the given initial condition at time ... confines itself to continuous-time time-invariant linear systems.
  19. [19]
    [PDF] University of Washington Lecture Notes for ME547 Linear Systems
    Feb 17, 2023 · ▷ time-invariant if its properties do not change with respect to time. ▷ Assuming the same initial conditions, if we shift u(t) by a constant ...
  20. [20]
    Basin of attraction - Scholarpedia
    Dec 29, 2016 · Its basin of attraction is the set of initial conditions leading to long-time behavior that approaches that attractor.
  21. [21]
  22. [22]
    [PDF] Predictability: Does the Flap of a Butterfly's Wings in Brazil Set Off a ...
    If the flap of a butterfly's wings can be instrumental in generating a tornado, it can equally well be instrumental in preventing a tornado. More generally, I ...
  23. [23]
    V. I. Oseledets, “A multiplicative ergodic theorem. Characteristic ...
    A multiplicative ergodic theorem. Characteristic Ljapunov, exponents of dynamical systems. Tr. Mosk. Mat. Obs., 19, MSU, M., 1968, 179–210.
  24. [24]
    Initial Conditions | Simulation Setup - SimScale
    May 14, 2025 · Initial conditions define the starting values for each solution field. Therefore, they play a vital role in the stability and computing time of ...
  25. [25]
    (PDF) The Significance of Initial Conditions in Simulations
    Aug 6, 2025 · Although initial conditions often significantly affect simulation results, little attention has been paid to test model sensitivity to them.
  26. [26]
    Spin‐up behavior and effects of initial conditions for an integrated ...
    Feb 26, 2015 · Initial conditions have significant, season-dependent effects on model outputs Dry initial conditions are more persistent than wet initial ...<|control11|><|separator|>
  27. [27]
    Assessing the role of initial conditions in the local structural ... - Nature
    Aug 19, 2021 · We present a systematic approach to detect problematic initial conditions of real-world systems biology models, that are usually not small.
  28. [28]
    Initial conditions for models of dynamical systems - ScienceDirect.com
    We consider the question of how to equip such a low-dimensional model with appropriate initial conditions, so that it faithfully reproduces the long-term ...Missing: definition | Show results with:definition
  29. [29]
    [PDF] Climate Models and Their Evaluation
    used an observational analysis for their initial conditions. The initial atmospheric component data are usually obtained from atmosphere-only integrations ...
  30. [30]
    The Value of Initial Condition Large Ensembles to Robust ...
    Aug 10, 2020 · Initial condition large ensembles of climate simulations have their scientific challenges, being expensive, possibly redundant, and biased ...
  31. [31]
    [PDF] Laplace's Demon and the Adventures of his Apprentices
    (3) Observational Omniscience: he is able to determine the true initial condition 0 x of the target system. If these conditions were met, the Demon would know ...
  32. [32]
    What Are the New Implications of Chaos for Unpredictability?
    It is widely believed and claimed by philosophers, mathematicians and physicists alike that chaos has a new implication for unpredictability, meaning that ...
  33. [33]
    [PDF] Determinism - PhilSci-Archive
    Jun 7, 2016 · A system is deterministic just in case the state of the system at one time fixes the state of the system at all future times.
  34. [34]
    [PDF] Laws, Initial Conditions and Physical Modality - PhilSci-Archive
    Apr 4, 2025 · In less formal language, initial conditions correspond to all possible values that can be attributed to the physical quantities in a dynamically.