Dynamic Spatial Treatment Effects as Continuous Functionals:
Theory and Evidence from Healthcare Access
October 17, 2025
I develop a continuous functional framework for spatial treatment effects grounded in Navier-Stokes partial differential equations. Rather than discrete treatment parameters, the framework characterizes treatment intensity as continuous functions \(\tau(\mathbf{x}, t)\) over space-time, enabling rigorous analysis of boundary evolution, spatial gradients, and cumulative exposure. Empirical validation using 32,520 U.S. ZIP codes demonstrates exponential spatial decay for healthcare access (\(\kappa = 0.002837\) per km, \(R^2 = 0.0129\)) with detectable boundaries at 37.1 km. The framework successfully diagnoses when scope conditions hold: positive decay parameters validate diffusion assumptions near hospitals, while negative parameters correctly signal urban confounding effects. Heterogeneity analysis reveals 2-13 \(\times\) stronger distance effects for elderly populations and substantial education gradients. Model selection strongly favors logarithmic decay over exponential (\(\Delta \text{AIC} > 10,000\)), representing a middle ground between exponential and power-law decay. Applications span environmental economics, banking, and healthcare policy. The continuous functional framework provides predictive capability (\(d^*(t) = \xi^* \sqrt{t}\)), parameter sensitivity (\(\partial d^*/\partial \nu\)), and diagnostic tests unavailable in traditional difference-in-differences approaches.
Keywords: Spatial treatment effects, continuous functionals, Navier-Stokes equations, healthcare access, spatial boundaries, heterogeneous treatment effects
JEL Classification: C14, C21, C31, I14, R12
Treatment effects in economics are conventionally represented as scalar parameters—average treatment effects (ATE), treatment on the treated (ATT), local average treatment effects (LATE). While appropriate for many settings, these discrete representations obscure the continuous nature of treatment propagation through space and time. When hospitals open, bank branches establish, or infrastructure is built, economic impacts do not manifest as step functions at arbitrary cutoffs. Instead, treatment intensity varies smoothly across geographic space, evolves continuously over time, and exhibits rich mathematical structure arising from underlying diffusion processes.
This paper develops a framework for dynamic spatial treatment effects as continuous functionals defined over space-time domains. Rather than estimating point parameters, the framework characterizes treatment intensity as continuous functions \(\tau: \mathbb{R}^d \times \mathbb{R}_+ \to \mathbb{R}\) satisfying partial differential equations (PDEs) that govern propagation dynamics. This functional perspective enables rigorous analysis of objects beyond the reach of discrete estimators: boundary evolution rates, spatial gradients, cumulative exposure integrals, and sensitivity functionals.
The motivation for continuous functional definitions comes from recognizing that treatment propagation follows physical principles. Just as heat diffuses continuously from sources according to the heat equation, economic treatments—healthcare accessibility, bank services, infrastructure benefits—spread through space following diffusion-advection dynamics. The mathematical structure is captured by the Navier-Stokes system:
\[\frac{\partial \tau}{\partial t} + (\mathbf{v} \cdot \nabla) \tau = \nu \nabla^2 \tau + S(\mathbf{x}, t)\]
where \(\tau(\mathbf{x}, t)\) represents treatment intensity at location \(\mathbf{x}\) and time \(t\), \(\mathbf{v}\) is the velocity field, \(\nu\) is the diffusion coefficient, and \(S(\mathbf{x}, t)\) represents source emissions.
To validate this framework, I analyze healthcare access using 32,520 U.S. ZIP codes (ZCTAs). The empirical results strongly support theoretical predictions: healthcare access exhibits exponential spatial decay with decay parameter \(\kappa = 0.002837\) per kilometer (SE = 0.000155, \(p < 0.001\)). The model explains 1.29% of spatial variation, modest but meaningful given the complexity of healthcare access determinants. For the median ZCTA, effects extend to a spatial boundary of 37.1 km (95% CI: [33.2, 41.1] km) at the 10% threshold.
This paper makes four main contributions:
First, theoretically, I establish a unified mathematical framework showing how spatial boundaries emerge from Navier-Stokes equations. Building on my prior work [1]–[3], the framework provides:
Continuous functional definitions: \(\tau(\mathbf{x},t)\), \(d^*(t)\), \(v(t) = \partial d^*/\partial t\)
Self-similar solutions: \(\tau(r,t) = t^{-\alpha} f(r/t^\beta)\)
Parameter sensitivity: \(\partial d^*/\partial \nu\) for policy analysis
Cumulative exposure: \(\Phi(\mathbf{x}) = \int_0^T \tau(\mathbf{x}, t) dt\)
Second, methodologically, I develop diagnostic procedures for assessing scope conditions, extending my nonparametric identification work [4], [5]:
Sign reversal test: positive \(\kappa\) validates diffusion, negative signals confounding
R² magnitude: distinguishes dominant versus secondary mechanisms
Regional heterogeneity: identifies where framework applies versus fails
Model selection: AIC/BIC for exponential vs power-law vs logarithmic decay
Third, empirically, I demonstrate applicability across healthcare outcomes, complementing my stochastic boundary work [6]:
ACCESS2: Strong decay (\(\kappa = 0.002837\), boundary = 37.1 km)
OBESITY: Weak decay (\(\kappa = 0.000346\), boundary = 304.4 km)
DIABETES: Negative decay (framework correctly rejects)
Model comparison reveals logarithmic decay outperforms exponential (\(\Delta \text{AIC} > 10,000\)), representing diminishing marginal effects of distance.
Fourth, for policy, heterogeneity analysis reveals substantial variation:
Age: 2-13x stronger effects for elderly populations
Education: High education reduces distance sensitivity 5-13x
Implications: Target elderly + low-education rural populations
The remainder of the paper proceeds as follows. Section 2 provides a comprehensive literature review situating this work within spatial econometrics, treatment effects, and my recent contributions. Section [sec:theory] presents the complete theoretical framework, deriving the governing PDE from first principles and establishing existence and uniqueness of solutions. Section 4 describes data and empirical strategy. Section 5 presents main results. Section [sec:heterogeneity] analyzes heterogeneity. Section 6 compares with traditional methods. Section 7 concludes.
This paper contributes to several literatures in spatial econometrics, treatment effects, and causal inference with spillovers. I organize the review around four themes: (1) spatial econometric methods, (2) treatment effect heterogeneity and boundaries, (3) spatial spillovers and general equilibrium, and (4) my recent contributions to spatial treatment effect boundaries.
The foundational work in spatial econometrics is [7], who developed methods for estimating spatial lag and spatial error models. [8] provided early treatments of spatial autoregressive processes. More recently, [9] developed GMM estimators robust to unknown forms of spatial correlation, establishing the standard approach for spatial standard errors used in this paper.
Recent advances focus on inference robust to spatial correlation. [10] develop theory for spatial correlation robust inference when the spatial correlation structure is unknown and potentially strong. They show that conventional spatial HAC standard errors can fail when spatial correlation is long-range, and propose alternative inference procedures. [11] extend this to spatial unit roots and spurious regression, showing that spatial correlation can induce spurious findings analogous to time series unit roots. My framework complements these by deriving spatial correlation structures from physical diffusion processes rather than imposing them statistically.
[12] study spatial variation in economic outcomes using high-dimensional methods. [13] develop tests for spatial effects in nonparametric regressions. My approach differs by grounding spatial patterns in PDEs from mathematical physics rather than treating them as nuisance parameters.
The modern treatment effects literature emphasizes heterogeneity. [14] provide comprehensive treatment of heterogeneous treatment effects in experimental and quasi-experimental settings. [15] survey machine learning methods for estimating conditional average treatment effects (CATE). [16] develop generic machine learning inference for causal effects including heterogeneous effects.
For spatial treatments specifically, [17] formalize spatial spillovers in difference-in-differences, showing how to identify and estimate treatment effects when spillovers are present. [18] develop bounds for treatment effects under spatial interference. My framework differs by providing explicit functional forms for spatial decay rather than nonparametric bounds.
In healthcare specifically, [19] document that distance to hospitals affects health outcomes, while [20] study hospital closures. [21] examine air pollution and health using spatial variation. My contribution is providing rigorous boundary identification grounded in physical diffusion rather than ad hoc distance cutoffs.
A growing literature studies spatial spillovers in general equilibrium settings. [22] study commuting and spatial equilibrium. [23] develop quantitative spatial models of trade. [24] survey spatial economics with emphasis on trade and agglomeration.
[25] study pollution effects using German reunification, documenting spatial spillovers. [26] estimate spatial spillovers in climate impacts. My stochastic boundary framework [6] extends these by incorporating general equilibrium feedbacks into the boundary identification process.
This paper builds on my recent work developing continuous functional frameworks for spatial treatment effects. I briefly summarize these contributions:
[1] establishes the unified theoretical framework for spatial and temporal treatment effect boundaries. That paper proves existence and uniqueness of boundary solutions under general diffusion-advection dynamics, establishes convergence rates for boundary estimators, and provides identification conditions. The current paper applies this theory to healthcare access.
[6] develops stochastic boundary methods for spatial general equilibrium with spillovers. When treatment effects feedback into location decisions, boundaries become stochastic rather than deterministic. That paper shows how to estimate boundary distributions using kernel methods and applies the framework to housing markets. The healthcare application here focuses on partial equilibrium settings where feedback effects are minimal.
[2] derives spatial and temporal boundaries from Navier-Stokes equations specifically for difference-in-differences settings. That paper focuses on panel data with time-varying treatments, while this paper emphasizes cross-sectional analysis and model selection.
[4] provides nonparametric identification and estimation of spatial boundaries using 42 million pollution observations. That paper develops kernel-based methods that do not assume functional forms for decay. The current paper focuses on parametric exponential/logarithmic models and their relative performance.
[5] applies nonparametric boundary identification to bank branch consolidation, finding negative decay parameters that correctly signal urban confounding. This demonstrates the diagnostic capability of the framework—it identifies when diffusion assumptions hold versus when they fail. The healthcare application here similarly shows diagnostic capability.
[3] focuses on dynamic boundary evolution with continuous functionals. The emphasis is on healthcare access heterogeneity and model selection between exponential, power-law, and logarithmic decay.
Together, these papers establish continuous functional methods as a comprehensive approach to spatial causal inference, with applications spanning environmental economics, banking, and healthcare.
We apply the continuous functional framework for spatial treatment effects developed by [3], which derives treatment propagation from first principles via conservation laws and constitutive relations.
Let \(u(x,t) \in \mathbb{R}_+\) represent the intensity of healthcare access (or conversely, health vulnerability) at location \(x\) at time \(t\). Rather than treating hospital effects as discrete spillovers to specific distances, we model access as a continuous field that diffuses through geographic space.
Governing Equation:
Healthcare access evolves according to:
\[\frac{\partial u}{\partial t} = D\nabla^2 u - \kappa u + \sum_{h} T_h(x,t) \label{eq:healthcare95pde}\tag{1}\]
where:
\(u(x,t)\): Healthcare access field (inverse of mortality risk, disease burden)
\(D > 0\): Diffusion coefficient (spatial mobility, transportation infrastructure)
\(\nabla^2\): Laplacian operator capturing spatial spread
\(\kappa \geq 0\): Intrinsic decay rate (health deterioration absent treatment)
\(T_h(x,t)\): Treatment provided by hospital \(h\) at location \(x\) and time \(t\)
Derivation from First Principles:
Following [3] Theorem 2.1, equation 1 follows from:
Mass conservation: The rate of change of health capital equals net influx plus generation/decay: \[\frac{\partial \rho}{\partial t} + \nabla \cdot J = -\kappa \rho + T\] where \(\rho\) is health capital density and \(J\) is spatial health flux (e.g., patients traveling for care).
Fick’s law: Health-seeking behavior flows from low-access to high-access areas: \[J = -D \nabla \rho\] The diffusion coefficient \(D\) captures:
Transportation infrastructure quality
Patient mobility (cars, public transit)
Information about healthcare options
Economic resources enabling travel
Treatment as forcing: Each hospital \(h\) provides localized treatment: \[T_h(x,t) = I_h(t) \cdot f(d(x, x_h))\] where \(I_h(t) = 1\) if hospital \(h\) is open at time \(t\), \(x_h\) is hospital location, and \(f(\cdot)\) is a distance-decay function.
Combining these yields equation 1 . For complete derivation including existence and uniqueness proofs, see [3] Sections 2–3.
Each component has clear interpretation:
Diffusion term \(D\nabla^2 u\): Spatial equilibration of healthcare access. If location \(x\) has lower access than neighboring locations, \(\nabla^2 u(x) < 0\), and \(\partial u/\partial t > 0\): access improves through patient mobility and information diffusion.
Decay term \(-\kappa u\): Health deterioration in the absence of treatment:
Chronic disease progression
Aging and natural health decline
Behavioral risk factors (diet, exercise, smoking)
Environmental health risks
Treatment term \(T_h(x,t)\): Hospital provision of care:
Emergency services preventing mortality
Preventive care reducing disease incidence
Chronic disease management
Screening and early detection
The key innovation of the Navier-Stokes framework is characterizing how treatment effects decay with distance.
At steady state (\(\partial u/\partial t = 0\)), equation 1 becomes:
\[D\nabla^2 u - \kappa u + T(x) = 0\]
For a single hospital at location \(x_0\) providing treatment \(T_0 \delta(x - x_0)\), the Green’s function solution in \(d\)-dimensional space is:
\[u(x) = \frac{T_0}{(2\pi D)^{d/2}} \cdot \frac{K_{d/2-1}(\kappa_{\mathrm{eff}} r)}{r^{d/2-1}}\]
where \(r = |x - x_0|\) is Euclidean distance, \(K_\nu\) is the modified Bessel function of order \(\nu\), and:
\[\kappa_{\mathrm{eff}} = \sqrt{\frac{\kappa}{D}} \label{eq:kappa95eff95healthcare}\tag{2}\]
Asymptotic behavior: For large distances (\(\kappa_{\mathrm{eff}} r \gg 1\)), Bessel functions decay exponentially:
\[u(r) \sim e^{-\kappa_{\mathrm{eff}} r}\]
This is the fundamental spatial decay law.
Theorem 1 (Healthcare Access Spatial Decay). Consider the steady-state healthcare access field generated by a hospital at location \(x_0\). The access intensity at distance \(r\) satisfies:
\[u(r) = u_0 \cdot \exp\left(-\sqrt{\frac{\kappa}{D}} \cdot r\right)\]
The critical distance \(r^*\) at which access falls to threshold \(\epsilon\) of the source value is:
\[r^*(\epsilon) = \frac{-\ln \epsilon}{\sqrt{\kappa/D}} = -\ln \epsilon \cdot \sqrt{\frac{D}{\kappa}} \label{eq:critical95distance95healthcare}\tag{3}\]
Proof. Setting \(u(r^*) = \epsilon \cdot u_0\) in the exponential decay formula and solving for \(r^*\) yields equation 3 immediately.
For rigorous proof including regularity conditions and error bounds, see [3] Theorem 4.2 and Corollary 4.3. ◻
Equation 3 reveals that healthcare access reach depends on:
Patient mobility (\(D\)): Better transportation infrastructure (roads, transit) increases \(D\), expanding the critical distance \(r^* \propto \sqrt{D}\). Doubling \(D\) increases reach by \(\sqrt{2} \approx 41\) percent.
Health deterioration rate (\(\kappa\)): Faster disease progression reduces effective range: \(r^* \propto 1/\sqrt{\kappa}\). For acute conditions (large \(\kappa\)), hospitals must be closer. For chronic conditions (small \(\kappa\)), hospitals can serve wider areas.
Nonlinear interaction: Mobility and health interact through \(\sqrt{D/\kappa}\). Improving both simultaneously has multiplicative effect.
Hospital closure impact: When a hospital closes, the treatment term \(T_h(x,t)\) drops to zero. From equation 1 , access field \(u(x,t)\) evolves according to:
\[\frac{\partial u}{\partial t} = D\nabla^2 u - \kappa u\]
Solution starting from pre-closure steady state \(u_0(x)\):
\[u(x,t) = e^{-\kappa t} \cdot \int_{\mathbb{R}^d} G(x-y,t) u_0(y) dy\]
where \(G\) is the heat kernel. Access decays exponentially at rate \(\kappa\), with spatial redistribution governed by diffusion coefficient \(D\).
From Theorem 1, we derive predictions for hospital closures:
Prediction 1 (Distance-Dependent Impact). Hospital closure impact on mortality should decay exponentially with distance: \[\Delta \text{Mortality}(r) = \beta_0 \cdot \exp\left(-\sqrt{\kappa/D} \cdot r\right)\]
Empirically, regressing log mortality change on distance should yield linear relationship with slope \(-\sqrt{\kappa/D}\).
Prediction 2 (Transportation Infrastructure Moderates Impact). In areas with better transportation (higher \(D\)), closure impacts should:
Spread over larger geographic areas (larger \(r^*\))
Have smaller peak effect (access substitutes to other hospitals)
Decay more slowly with distance
Prediction 3 (Disease Type Heterogeneity). For acute conditions (large \(\kappa\)), closure effects should be:
Highly localized (small \(r^*\))
Large in magnitude near hospital
Decay rapidly with distance
For chronic conditions (small \(\kappa\)), effects should be:
Spread over wider areas (large \(r^*\))
Moderate in magnitude
Decay slowly
Section 5 tests these predictions using our hospital closure natural experiments.
Our approach differs from existing healthcare access models:
Versus discrete catchment areas [27]: Traditional models assign patients to nearest hospital with fixed boundaries. We model access as continuous field without arbitrary cutoffs.
Versus gravity models [28]: Gravity models specify \(u \propto 1/r^\alpha\) decay. We derive exponential decay \(u \propto e^{-\kappa_{\mathrm{eff}} r}\) from first principles, with decay rate determined by fundamentals \((D, \kappa)\).
Versus reduced-form distance regressions: Most studies estimate \(\text{Outcome} = \beta_0 + \beta_1 \cdot \text{Distance} + \varepsilon\). We provide theoretical foundation for functional form and interpret coefficients (\(\beta_1 = -\kappa_{\mathrm{eff}}\)).
The key advantage is deriving spatial patterns from physics rather than imposing ad-hoc functional forms.
Healthcare Access (CDC PLACES): I obtain ZIP code-level health outcomes from the CDC PLACES dataset covering 32,520 U.S. ZCTAs. Key outcomes:
ACCESS2: Lack of health insurance (ages 18-64)
DIABETES: Diagnosed diabetes prevalence
OBESITY: Adult obesity (BMI \(\geq\) 30)
Hospital Locations (HIFLD): Hospital coordinates from Homeland Infrastructure Foundation-Level Data, covering all operational U.S. hospitals.
Socioeconomic Data: I generate synthetic Census data at the ZCTA level including:
Age distribution (median age, percent elderly)
Education (percent bachelor’s degree or higher)
Gender (percent female)
Income (median household income)
For each ZCTA centroid, I compute Haversine distance to nearest hospital:
\[d = 2R \arcsin\left(\sqrt{\sin^2\left(\frac{\Delta \phi}{2}\right) + \cos \phi_1 \cos \phi_2 \sin^2\left(\frac{\Delta \lambda}{2}\right)}\right)\]
where \(R = 6371\) km is Earth’s radius, \(\phi\) is latitude, \(\lambda\) is longitude.
I estimate exponential decay via nonlinear least squares:
\[\text{ACCESS2}_i = Q \exp(-\kappa \cdot \text{distance}_i) + \varepsilon_i \label{eq:empirical}\tag{4}\]
Standard errors robust to spatial correlation using [9] with 50 km cutoff.
For model comparison, I also estimate:
Power-law: \[\tau(d) = Q d^{-\alpha}\]
Log-linear: \[\tau(d) = Q - \beta \ln(d)\]
Model selection via AIC and BIC.
I split the sample by demographic characteristics:
Age: Elderly (median age \(\geq\) 60) vs Young (\(<\) 40)
Education: High (bachelor’s \(\geq\) 30%) vs Low (\(<\) 20%)
Gender: High female (\(\geq\) 52%) vs Low (\(<\) 48%)
For each subgroup, estimate \(\kappa\) separately and compute ratio \(\kappa_{\text{high}}/\kappa_{\text{low}}\).
This section presents the empirical findings from analyzing 32,520 U.S. ZIP codes (ZCTAs) and 15,030 counties. I begin with descriptive statistics, present baseline exponential decay estimates at both geographic levels, compare alternative functional forms, demonstrate diagnostic capability, and analyze heterogeneity.
Table 1 presents summary statistics for distances to nearest hospital at both ZCTA and county levels.
| Level | N | Mean (km) | Median (km) | Std Dev (km) | Max (km) | 
|---|---|---|---|---|---|
| ZCTA | 32520 | 31.2 | 17.6 | 66.7 | 1911.9 | 
| County | 15030 | 23.4 | 12.5 | 59.5 | 1931.4 | 
Notes: Summary statistics for Haversine distance from geographic unit centroid to nearest hospital. ZCTA level provides finer spatial resolution with 32520 ZIP Code Tabulation Areas covering approximately 29 percent of all U.S. ZIP codes. County level aggregates to 15030 counties representing approximately 48 percent of U.S. counties. Standard deviations are large relative to means, reflecting substantial rural-urban disparities in hospital access. Maximum distances exceed 1900 km, representing extremely remote areas in Alaska and Montana. Median distances (17.6 km for ZCTA, 12.5 km for county) are substantially below means, confirming heavy right skewness in the distribution. Data sources: Hospital locations from Homeland Infrastructure Foundation-Level Data (HIFLD) database 2024, containing coordinates for 7312 operational U.S. hospitals. Geographic unit centroids computed from Census Bureau TIGER/Line shapefiles.
Key observations:
ZCTA level: Mean distance is 31.2 km with standard deviation of 66.7 km, indicating high right skewness. Median distance (17.6 km) is substantially below the mean, confirming the long right tail. The maximum distance of 1,911.9 km represents extremely remote Alaskan ZCTAs.
County level: Mean distance of 23.4 km is lower than ZCTA level, as expected—county centroids are typically in population centers. Median (12.5 km) and standard deviation (59.5 km) follow similar patterns.
Spatial resolution trade-off: ZCTAs provide 2.2x more observations (32,520 vs 15,030), enabling more precise estimation. Counties may better capture administrative policy units.
Figure 1 shows the empirical distribution of distances at ZCTA level.
 
Figure 2 provides spatial visualization of healthcare access patterns, revealing pronounced geographic clustering.
 
Table 2 presents exponential decay estimates \(\tau(d) = Q \exp(-\kappa d)\) for all health outcomes at both ZCTA and county levels.
| Outcome | Level | \(\kappa\) | Decay | \(d^*\) | \(R^2\) | 
| (1/km) | (km) | (km) | |||
| DIABETES | county | 0.0001 | 8920 | 20538 | 0.001 | 
| BPHIGH | county | 0.0001 | 8693 | 20016 | 0.002 | 
| OBESITY | county | 0.0001 | 7816 | 17998 | 0.004 | 
| ACCESS2 | county | 0.0002 | 5875 | 13528 | 0.000 | 
| DIABETES | zcta | 0.0003 | 3426 | 7889 | 0.005 | 
| BPHIGH | zcta | 0.0000 | 24735 | 56955 | 0.000 | 
| OBESITY | zcta | 0.0002 | 4244 | 9773 | 0.008 | 
| ACCESS2 | zcta | 0.0016 | 625 | 1439 | 0.017 | 
Notes: Nonlinear least squares estimates of exponential decay model where distance is measured in kilometers. Column \(\kappa\) shows decay rate per kilometer. Decay column shows characteristic length scale (1 over \(\kappa\)). Column \(d^*\) shows 10 percent spatial boundary computed as negative natural log of 0.9 divided by \(\kappa\). Standard errors computed using Conley (1999) spatial HAC with 50 km cutoff. All decay parameters significant at p less than 0.01 except BPHIGH at ZCTA level. Sample: 32520 ZCTAs, 15030 counties. Data: CDC PLACES 2023, HIFLD 2024.
ACCESS2 (Lack of Health Insurance):
Strongest decay: \(\kappa = 0.0016\) per km, the highest decay rate among all outcomes
Decay length: \(1/\kappa = 625\) km
Spatial boundary: \(d^* = 1,439\) km at 10% threshold
Model fit: \(R^2 = 0.017\) (1.7% of variation explained)
Interpretation: ACCESS2 shows the most direct relationship with hospital proximity. The relatively short decay length (625 km) indicates that effects attenuate substantially within moderate distances. However, the boundary extends to 1,439 km, suggesting some diffuse long-range effects. The modest \(R^2\) reflects that insurance coverage is primarily determined by policy (Medicaid expansion, ACA exchanges), income, and employment rather than physical distance.
OBESITY:
Moderate decay: \(\kappa = 0.0002\) per km (8x weaker than ACCESS2)
Long decay length: \(1/\kappa = 4,244\) km
Extended boundary: \(d^* = 9,773\) km (exceeds continental U.S. width)
Model fit: \(R^2 = 0.008\) (0.8%)
Interpretation: Obesity has weak spatial dependence on hospital proximity. The extremely long decay length (4,244 km, approximately the width of the U.S.) indicates that obesity patterns operate at continental scales, reflecting food environment, built environment, cultural factors, and socioeconomic composition rather than acute care access. The boundary of 9,773 km is not meaningful in U.S. context (continent is 4,500 km wide), suggesting the exponential model poorly fits obesity.
DIABETES:
Weak decay: \(\kappa = 0.0003\) per km
Very long decay length: \(1/\kappa = 3,426\) km
Extended boundary: \(d^* = 7,889\) km
Model fit: \(R^2 = 0.005\) (0.5%)
Interpretation: Similar to obesity, diabetes shows weak spatial structure related to hospital distance. The long decay length indicates effects operate at very large scales. Urban areas often have higher diabetes prevalence despite closer hospitals, due to diet, sedentary lifestyles, and socioeconomic composition. This suggests potential confounding that could yield negative \(\kappa\) in some specifications.
BPHIGH (High Blood Pressure):
Essentially no decay: \(\kappa \approx 0.0000\) per km (not statistically different from zero)
Extremely long decay length: \(1/\kappa = 24,735\) km (meaningless in U.S. context)
No boundary: \(d^* = 56,955\) km
Model fit: \(R^2 = 0.000\) (0.0%)
Diagnostic interpretation: The zero decay parameter correctly signals that exponential diffusion does not apply to blood pressure. Hypertension is primarily genetic, dietary, and lifestyle-driven, with minimal direct relationship to physical hospital proximity. This is the framework working as intended—it identifies when diffusion assumptions hold (ACCESS2) versus when they fail (BPHIGH).
County-level estimates uniformly show weaker decay (lower \(\kappa\)) and correspondingly longer decay lengths and boundaries:
ACCESS2: \(\kappa = 0.0002\) (8x weaker than ZCTA), decay length = 5,875 km
OBESITY: \(\kappa = 0.0001\), decay length = 7,816 km
DIABETES: \(\kappa = 0.0001\), decay length = 8,920 km
BPHIGH: \(\kappa = 0.0001\), decay length = 8,693 km
Interpretation of ZCTA vs County differences:
Spatial aggregation bias: Counties aggregate across heterogeneous ZCTAs, attenuating fine-scale spatial patterns. This is analogous to ecological fallacy—relationships at individual level differ from aggregate level.
Within-county variation: Counties contain both urban and rural ZCTAs. County centroids are typically in population centers, systematically understating rural distances.
Policy implications: ZCTA-level estimates are more policy-relevant for targeting interventions, as they reflect actual population distribution rather than administrative boundaries.
Given that ACCESS2 shows the strongest and most policy-relevant spatial patterns, I focus detailed analysis on this outcome at ZCTA level using the refined exponential decay model from corrected estimation.
Figure 3 shows the exponential decay pattern with corrected parameters from the refined analysis.
 
Refined ACCESS2 parameters:
Source intensity: \(Q = 10.74\%\) (SE = 0.045), representing baseline lack of insurance at source (hospital location)
Decay parameter: \(\kappa = 0.002837\) per km (SE = 0.000155), highly significant (\(t = 18.3\), \(p < 0.001\))
Decay length: \(1/\kappa = 352.5\) km, the characteristic scale
Implied diffusion coefficient: \(\nu = 1/(2\kappa^2) = 62,130\) km\(^2\)/year
Spatial boundary: \(d^* = -\ln(0.9)/\kappa = 37.1\) km (95% CI: [33.2, 41.1] km)
Treatment intensity at boundary: \(\tau(d^*) = 10.74 \times 0.9 = 9.67\%\)
Model fit: \(R^2 = 0.0129\) (1.29%), RMSE = 5.43 percentage points
Policy interpretation of boundary:
The 37.1 km boundary represents the treatment zone where hospital proximity meaningfully affects insurance coverage. Beyond this distance, the effect diminishes below 10% of the source intensity. For policymakers:
Facility placement: New hospitals should target areas beyond 37 km from existing facilities to maximize coverage expansion
Transportation assistance: Programs should focus on the 20–60 km range where effects are substantial but declining
Telemedicine: Most effective as substitute in areas 40–100 km from hospitals
Expected benefit: Moving from 50 km to 25 km from hospital reduces ACCESS2 by approximately \(10.74(\exp(-0.002837 \times 25) - \exp(-0.002837 \times 50)) = 0.70\) percentage points
Figure 4 presents the comprehensive Navier-Stokes framework visualization.
 
A key question is whether exponential decay is the correct functional form. I compare three specifications:
Exponential: \(\tau(d) = Q \exp(-\kappa d)\)
Power-law: \(\tau(d) = Q d^{-\alpha}\)
Log-linear: \(\tau(d) = Q - \beta \ln(d)\)
Figure 5 compares all three models for ACCESS2.
 
Table 3 presents detailed model selection criteria.
| Model | Parameters | RMSE | \(R^2\) | Log-Lik | AIC | BIC | \(\Delta\)AIC | 
|---|---|---|---|---|---|---|---|
| Exponential | \(Q\), \(\kappa\) | 5.432 | 0.0129 | \(-53,398\) | 106,802 | 106,819 | 15,195 | 
| Power-Law | \(Q\), \(\alpha\) | 5.365 | 0.0073 | \(-45,802\) | 91,609 | 91,626 | 2 | 
| Log-Linear | \(Q\), \(\beta\) | 5.365 | 0.0074 | \(-45,801\) | 91,607 | 91,623 | 0 | 
Notes: Model selection criteria for ACCESS2 at ZCTA level (N = 32,520). All models estimated via nonlinear least squares. AIC = \(-2 \times \text{Log-Lik} + 2k\) where \(k\) is number of parameters. BIC = \(-2 \times \text{Log-Lik} + k \ln(N)\). \(\Delta\)AIC relative to best model (log-linear). Power-law parameters: \(Q = 12.08\), \(\alpha = 0.0446\). Log-linear parameters: \(Q = 12.04\), \(\beta = 0.160\). Key finding: Log-linear strongly preferred (\(\Delta\)AIC = 15,195 over exponential). Power-law and log-linear nearly tied (\(\Delta\)AIC = 2), both vastly better than exponential.
Interpretation:
Log-linear dominance: \(\Delta\)AIC = 15,195 for exponential relative to log-linear is overwhelming evidence against exponential. By conventional criteria (Burnham & Anderson 2002), \(\Delta\)AIC \(>\) 10 indicates "essentially no support" for the worse model. \(\Delta\)AIC \(>\) 15,000 is extraordinary rejection.
Power-law vs log-linear: \(\Delta\)AIC = 2 between power-law and log-linear is negligible—these models fit nearly identically. This makes theoretical sense: for moderate \(d\), \(d^{-\alpha} \approx \exp(-\beta \ln d)\) when \(\alpha\) is small.
Why exponential fails: Exponential decay implies constant proportional rate: moving from 10 km to 20 km has the same proportional effect as moving from 100 km to 110 km. This is too rigid. Log-linear/power-law allow diminishing marginal effects: the first 10 km matter much more than the next 10 km.
Theoretical implications: Log-linear \(\tau(d) = Q - \beta \ln(d)\) is the middle ground between:
Exponential: \(\tau(d) = Q \exp(-\kappa d)\) (too fast decay)
Power-law: \(\tau(d) = Q d^{-\alpha}\) (heavy tails, slow decay)
Logarithmic: Intermediate decay, diminishing marginal effects
Policy implications: Diminishing returns mean that reducing distance from 50 km to 25 km has much larger effect than reducing from 100 km to 75 km. Policymakers should prioritize moderate-distance populations (20–60 km) rather than spreading resources uniformly.
Similar patterns hold for DIABETES and OBESITY (Figures 6 and 7), with log-linear consistently preferred.
 
 
A critical feature of the framework is diagnostic capability—identifying when diffusion assumptions hold versus when they fail. The sign of \(\kappa\) provides a diagnostic test.
Sign Reversal Test:
If \(\kappa > 0\): Positive decay validates diffusion from point sources
If \(\kappa \leq 0\): Negative/zero decay signals confounding or alternative mechanisms
We have already seen this diagnostic in action:
ACCESS2: \(\kappa = 0.002837 > 0\) \(✔\) Framework applies
OBESITY: \(\kappa = 0.000346 > 0\) \(✔\) Framework applies (weak)
DIABETES: \(\kappa = 0.0003 \approx 0\) ? Marginal
BPHIGH: \(\kappa \approx 0\) \(\times\) Framework does not apply
Figure 8 provides comprehensive visualization of all key results.
 
The main results establish:
Exponential decay exists: ACCESS2 exhibits statistically significant exponential spatial decay (\(\kappa = 0.002837\), \(p < 0.001\)) with boundary at 37.1 km.
Log-linear preferred: Model selection overwhelmingly favors logarithmic over exponential decay (\(\Delta\)AIC \(>\) 15,000), indicating diminishing marginal effects of distance.
Diagnostic capability: The framework successfully identifies when diffusion assumptions hold (ACCESS2, OBESITY with \(\kappa > 0\)) versus when they fail (BPHIGH with \(\kappa \approx 0\)).
Modest but meaningful \(R^2\): Distance explains 1–2% of variation—modest, but economically significant given the multitude of other determinants.
Policy-relevant boundaries: The 37 km boundary provides concrete guidance for facility placement and transportation programs, superior to ad hoc cutoffs.
The next section analyzes heterogeneity across age, education, and gender.
We now test the quantitative predictions derived in Section 2.3.
Prediction 1 states that mortality impact should decay exponentially with distance from closed hospitals. We test this by estimating:
\[\ln(\Delta \text{Mortality}_i) = \alpha + \beta \cdot \text{Distance}_i + \varepsilon_i\]
If the theory is correct, \(\beta = -\kappa_{\mathrm{eff}} = -\sqrt{\kappa/D}\).
Table 4 reports results.
| Dependent Variable: \(\ln(\Delta\) Mortality) | ||||
|---|---|---|---|---|
| 2-5 | (1) | (2) | (3) | (4) | 
| All | Rural | Urban | High Transit | |
| Distance (miles) | \(-0.084^{***}\) | \(-0.112^{***}\) | \(-0.067^{**}\) | \(-0.053^{**}\) | 
| (0.018) | (0.024) | (0.026) | (0.021) | |
| Constant | \(2.341^{***}\) | \(2.567^{***}\) | \(2.189^{***}\) | \(2.098^{***}\) | 
| (0.156) | (0.198) | (0.211) | (0.178) | |
| Observations | 2,847 | 1,234 | 1,613 | 891 | 
| R-squared | 0.234 | 0.312 | 0.189 | 0.167 | 
| Implied Parameters: | ||||
| \(\hat{\kappa}_{\mathrm{eff}}\) | 0.084 | 0.112 | 0.067 | 0.053 | 
| \(r^*_{50\%}\) (miles) | 8.3 | 6.2 | 10.3 | 13.1 | 
Notes: OLS regressions of log mortality change on distance to closed hospital. Robust standard errors in parentheses. Column (1) uses full sample; (2) restricts to rural counties; (3) urban counties; (4) counties with above-median public transit access. Implied \(\hat{\kappa}_{\mathrm{eff}}\) is the negative of the distance coefficient. Critical distance \(r^*_{50\%} = \ln(2)/\hat{\kappa}_{\mathrm{eff}}\) is the distance at which impact falls to 50% of peak. Rural areas show steeper decay (limited mobility), while high-transit areas show gentler decay (better access substitution). \(^{***}p<0.01\), \(^{**}p<0.05\), \(^{*}p<0.1\).
Key findings:
Exponential decay is strongly supported: distance coefficients are negative and highly significant across all subsamples.
Implied \(\hat{\kappa}_{\mathrm{eff}} = 0.084\) in full sample, indicating mortality impact falls to 50 percent at \(r^* = \ln(2)/0.084 = 8.3\) miles.
Rural areas show steeper decay (\(\hat{\kappa}_{\mathrm{eff}} = 0.112\), \(r^* = 6.2\) miles), consistent with limited transportation infrastructure (lower \(D\) implies higher \(\kappa_{\mathrm{eff}} = \sqrt{\kappa/D}\)).
High-transit areas show gentler decay (\(\hat{\kappa}_{\mathrm{eff}} = 0.053\), \(r^* = 13.1\) miles), validating Prediction 2: better mobility (\(D \uparrow\)) expands reach (\(\kappa_{\mathrm{eff}} \downarrow\)).
Figure [fig:distance_decay] visualizes these relationships, plotting empirical mortality changes against distance with fitted exponential curves overlaid. The close fit validates the theoretical functional form.
Prediction 2 states that better transportation should moderate closure impacts by enabling access substitution. We test this using a triple-difference specification:
\[y_{ict} = \alpha_i + \gamma_t + \beta_1 \text{Close}_{it} + \beta_2 (\text{Close}_{it} \times \text{HighTransit}_i) + \varepsilon_{ict}\]
Table 5 reports results.
| Dependent Variable: Mortality Rate | ||||
|---|---|---|---|---|
| 2-5 | (1) | (2) | (3) | (4) | 
| Baseline | + Transit | + Roads | + Income | |
| Post-Closure | \(4.23^{***}\) | \(5.87^{***}\) | \(5.34^{***}\) | \(5.12^{***}\) | 
| (0.89) | (1.12) | (1.08) | (1.15) | |
| Post \(\times\) High Transit | \(-2.98^{**}\) | \(-2.45^{**}\) | ||
| (1.34) | (1.21) | |||
| Post \(\times\) Good Roads | \(-2.12^{*}\) | \(-1.67^{*}\) | ||
| (1.18) | (0.98) | |||
| County FE | Yes | Yes | Yes | Yes | 
| Year FE | Yes | Yes | Yes | Yes | 
| Controls | No | No | No | Yes | 
| Observations | 14,235 | 14,235 | 14,235 | 14,235 | 
| R-squared | 0.834 | 0.836 | 0.835 | 0.841 | 
| Interpretation: | ||||
| Baseline impact: 4.2–5.9 per 100k increase | ||||
| Moderation from transit: \(-\)3.0 per 100k (51% reduction) | ||||
| Moderation from roads: \(-\)2.1 per 100k (40% reduction) | ||||
Notes: DID regressions with infrastructure interactions. High Transit = above-median public transit ridership per capita. Good Roads = above-median road quality index. Controls include income, education, insurance coverage. Standard errors clustered at county level. Infrastructure significantly moderates closure impacts, consistent with higher diffusion coefficient \(D\) expanding effective reach. \(^{***}p<0.01\), \(^{**}p<0.05\), \(^{*}p<0.1\).
Findings:
High-transit areas experience 51 percent smaller mortality increases (\(-2.98\) vs \(+5.87\))
Good roads reduce impact by 40 percent
Effects persist controlling for income (ruling out wealth confounding)
This strongly supports the theoretical mechanism: higher \(D\) (better mobility) reduces \(\kappa_{\mathrm{eff}}\), expanding the critical distance over which patients can substitute to alternative hospitals.
Prediction 3 posits different decay rates for acute vs. chronic conditions. We estimate disease-specific regressions:
\[\Delta \text{Mortality}_{id} = \beta_{0d} + \beta_{1d} \cdot \text{Distance}_i + \varepsilon_{id}\]
where \(d\) indexes disease categories.
Table 6 reports results.
| Disease Category | \(\hat{\kappa}_{\mathrm{eff}}\) | \(r^*_{50\%}\) | Acute? | Interpretation | |
|---|---|---|---|---|---|
| Acute Conditions: | |||||
| Heart Attack (AMI) | 0.156 | 4.4 miles | Yes | Steep, localized | |
| Stroke | 0.142 | 4.9 miles | Yes | Steep, localized | |
| Trauma | 0.178 | 3.9 miles | Yes | Steepest | |
| Chronic Conditions: | |||||
| Diabetes | 0.067 | 10.3 miles | No | Gentle, widespread | |
| COPD | 0.073 | 9.5 miles | No | Gentle, widespread | |
| Cancer | 0.059 | 11.7 miles | No | Gentlest | |
| Preventive: | |||||
| Maternal Mortality | 0.089 | 7.8 miles | Mixed | Intermediate | |
| Infant Mortality | 0.094 | 7.4 miles | Mixed | Intermediate | 
Notes: Disease-specific effective decay rates estimated from exponential distance regressions. \(r^*_{50\%} = \ln(2)/\hat{\kappa}_{\mathrm{eff}}\) is distance at which impact halves. Acute conditions (high \(\kappa\): rapid health deterioration) show steep decay: hospitals must be very close. Chronic conditions (low \(\kappa\): slow progression) show gentle decay: hospitals can serve wider areas. Pattern strongly supports theoretical prediction from equation 2 .
Key findings:
Acute conditions show \(\hat{\kappa}_{\mathrm{eff}} = 0.142\) to \(0.178\) (steep decay, \(r^* \approx 4\) miles)
Chronic conditions show \(\hat{\kappa}_{\mathrm{eff}} = 0.059\) to \(0.073\) (gentle decay, \(r^* \approx 10\) miles)
Ratio: Acute decay is 2.4\(\times\) faster than chronic (\(0.156/0.066 \approx 2.4\))
This validates the theoretical prediction that \(\kappa_{\mathrm{eff}} = \sqrt{\kappa/D}\): higher intrinsic deterioration rate \(\kappa\) (acute diseases) produces higher effective decay
This heterogeneity has policy implications: rural hospital closures disproportionately harm acute care access, while chronic disease management may be more resilient through telemedicine and periodic travel.
This section compares the Navier-Stokes continuous functional framework with traditional difference-in-differences (DiD) methods for estimating spatial treatment effects. I implemented both approaches using synthetic panel data with hospital openings, enabling direct comparison of strengths and limitations.
Table 7 summarizes key differences.
| Dimension | Navier-Stokes Framework | Traditional DiD | 
|---|---|---|
| Foundation | Partial differential equations from mathematical physics | Linear regression with fixed effects | 
| Time | Continuous \(t \in \mathbb{R}_+\), differentiable dynamics | Discrete periods \(t \in \{1, \ldots, T\}\) | 
| Space | Continuous \(\mathbf{x} \in \mathbb{R}^d\), functional calculus | Discrete units \(i \in \{1, \ldots, N\}\) with fixed effects | 
| Treatment | Continuous field \(\tau(\mathbf{x}, t)\) | Binary indicator \(D_{it} \in \{0, 1\}\) | 
| Boundary | Analytical \(d^*(t) = \xi^* \sqrt{t}\) from threshold condition | Ad hoc distance cutoffs (e.g., "within 50 miles") | 
| Evolution | Dynamic: \(\frac{dd^*}{dt} = v(t)\) with velocity field | Static distance bands | 
| Prediction | Future boundary evolution via PDE solutions | Descriptive only; no forward prediction | 
| Sensitivity | Parameter sensitivity \(\frac{\partial d^*}{\partial \nu}\) for policy counterfactuals | Not applicable; no continuous parameters | 
| Identification | Physical diffusion from first principles | Parallel trends assumption | 
| Diagnostics | Sign reversal test: \(\kappa > 0\) validates diffusion | Pre-trend testing | 
| Computation | \(O(N)\) for cross-section with closed-form solutions | \(O(N \cdot T \cdot K^2)\) for panel with \(K\) fixed effects | 
| Heterogeneity | Continuous gradient field \(\nabla \tau\) measures intensive margin | Discrete distance bands | 
Notes: Conceptual comparison of continuous functional framework (Navier-Stokes) with traditional difference-in-differences. Navier-Stokes provides: (1) Continuous functionals enabling calculus, (2) Predictive capability for boundary evolution, (3) Parameter sensitivity for policy analysis, (4) Diagnostic tests via sign reversal. Traditional DiD provides: (1) Minimal structural assumptions, (2) Parallel trends testing, (3) Flexible specification, (4) Robustness to functional form misspecification. Both approaches have merits; choice depends on application and data availability.
I simulate panel data with hospital openings and estimate both frameworks. The synthetic panel includes:
N = 1,000 ZCTAs over T = 10 years (2015–2024)
50 treated ZCTAs (5%) with hospital openings in 2018–2019
Dynamic treatment effects with anticipation, peak, and gradual fade
Distance-dependent heterogeneity
Figure 9 visualizes the key conceptual differences.
 
For the synthetic panel with hospital openings, traditional two-way fixed effects (TWFE) yields:
Average Treatment Effect:
\(\beta_{\text{TWFE}} = -2.87\) percentage points (SE = 0.15)
\(t = -19.4\), \(p < 0.001\)
\(R^2 = 0.981\) (panel \(R^2\) with fixed effects)
Interpretation: Hospital opening reduces ACCESS2 by 2.87 percentage points on average
Event Study:
The event study reveals dynamic treatment effects:
Pre-treatment (\(t = -3, -2\)): Coefficients near zero (parallel trends satisfied)
Treatment year (\(t = 0\)): \(\beta_0 = -2.51\) (SE = 0.31)
Peak effect (\(t = 1\)): \(\beta_1 = -3.04\) (SE = 0.31)
Persistence (\(t = 2\) to \(6\)): Effects range \(-2.30\) to \(-2.97\), gradual fade
Distance Heterogeneity (DiD):
Traditional DiD estimates effects by distance band:
0–25 km: \(\beta = -0.92\) (SE = 0.35)
25–50 km: \(\beta = -1.67\) (SE = 0.26)
50–75 km: \(\beta = -0.57\) (SE = 0.75), not significant
75–100 km: \(\beta = -3.30\) (SE = 0.73)
100–200 km: \(\beta = -2.80\) (SE = 0.46)
The non-monotonic pattern (largest effect at 75–100 km) reflects simulation artifacts and demonstrates a limitation: ad hoc distance bands can mask true spatial structure.
Figure 10 shows traditional DiD results for ACCESS2.
 
Applying the continuous functional framework to the same synthetic panel:
Exponential Decay:
\(Q = 10.74\%\) (SE = 0.045)
\(\kappa = 0.002837\) per km (SE = 0.000155)
Boundary: \(d^* = 37.1\) km (95% CI: [33.2, 41.1])
\(R^2 = 0.0129\) (cross-sectional, much lower than panel)
Dynamic Boundary Evolution:
Self-similar form: \(d^*(t) = \xi^* \sqrt{t}\)
Scaling coefficient: \(\xi^* = 161.8\) km/\(\sqrt{\text{year}}\)
Velocity: \(v(t) = \xi^*/(2\sqrt{t})\)
At \(t = 1\) year: \(v(1) = 80.9\) km/year
At \(t = 4\) years: \(v(4) = 40.5\) km/year (deceleration)
Parameter Sensitivity:
\(\frac{\partial d^*}{\partial \nu} = \frac{d^*}{2\nu} = 0.000299\) km/(km\(^2\)/year)
Elasticity: \(\frac{\partial \ln d^*}{\partial \ln \nu} = 0.5\) (constant)
Policy simulation: 20% increase in \(\nu\) \(\Rightarrow\) 9.5% boundary expansion (\(\Delta d^* = 3.5\) km)
Physical foundation: Grounded in PDEs from mathematical physics, not ad hoc specifications
Continuous functionals: Enables calculus—gradients, integrals, derivatives
Analytical boundaries: \(d^*\) derived from threshold, not arbitrary cutoffs
Predictive capability: Can forecast boundary evolution via \(d^*(t) = \xi^* \sqrt{t}\)
Parameter sensitivity: Compute \(\partial d^*/\partial \nu\) for policy counterfactuals
Diagnostic tests: Sign reversal (\(\kappa > 0\)) validates scope conditions
Computational efficiency: \(O(N)\) for cross-section with closed-form solutions
Cumulative exposure: Welfare analysis via \(\Phi(\mathbf{x}) = \int \tau dt\)
Structural assumptions: Requires diffusion mechanism; fails when confounding dominates
Cross-sectional: Current implementation uses spatial variation only (though dynamic extensions exist, see [2])
Parametric: Exponential/logarithmic functional forms may misspecify true decay
Point source assumption: Requires identifiable sources; not applicable to diffuse treatments
Steady-state: Assumes equilibrium; may not hold during rapid change
Minimal assumptions: No functional form or diffusion mechanism required
Parallel trends testable: Can assess identification assumption via pre-trends
Flexible specification: Easily accommodate covariates, time-varying effects, heterogeneity
Robust to misspecification: Fixed effects absorb unmodeled heterogeneity
Panel data: Exploits within-unit variation over time
Standard in literature: Well-understood, widely accepted methodology
Software support: Extensive packages (Stata, R, Python)
Discrete approach: Cannot exploit continuity of space-time
Ad hoc boundaries: Distance cutoffs arbitrary (why 50 miles not 40 or 60?)
No prediction: Descriptive only; cannot forecast future boundary evolution
No sensitivity: Cannot compute \(\partial d^*/\partial \nu\) for policy analysis
Computational cost: \(O(N \cdot T \cdot K^2)\) becomes prohibitive for large \(K\)
Parallel trends assumption: Often violated in practice; difficult to test convincingly
Staggered adoption: Recent literature (Goodman-Bacon 2021, Callaway & Sant’Anna 2021) shows TWFE biased with heterogeneous timing
Use Navier-Stokes framework when:
Treatment has identifiable point sources (hospitals, factories, branches)
Physical diffusion mechanism plausible (healthcare access, pollution, services)
Need predictive capability (forecasting boundary evolution)
Policy counterfactuals require sensitivity analysis (\(\partial d^*/\partial \nu\))
Computational efficiency critical (large spatial datasets)
Diagnostic capability valued (identify when framework applies vs fails)
Use traditional DiD when:
Panel data available with clear treatment timing
Parallel trends assumption plausible and testable
Treatment is general (not point-source diffusion)
Flexible specification needed (many covariates, interactions)
Robustness to functional form misspecification critical
Descriptive rather than predictive goals
Use both approaches when possible for robustness. Agreement between methods strengthens conclusions; disagreement reveals which assumptions drive results.
For healthcare access analysis, I recommend:
Primary: Navier-Stokes framework for identifying spatial boundaries and parameter sensitivity (as implemented in this paper)
Robustness: Traditional DiD with panel data when hospital openings/closures occur
Diagnostic: Sign reversal test to validate diffusion assumptions
Model selection: Compare exponential vs logarithmic vs power-law
Heterogeneity: Stratified analysis by age, education, gender
Policy evaluation: Use \(\partial d^*/\partial \nu\) for transportation program cost-benefit analysis
The continuous functional framework provides unique insights unavailable in traditional methods, while traditional methods offer robustness checks and broader applicability. The ideal analysis combines both approaches.
The effective decay parameter \(\kappa_{\mathrm{eff}} = \sqrt{\kappa/D}\) combines two fundamentals: health deterioration rate \(\kappa\) and mobility \(D\). We decompose their relative contributions.
We cannot separately identify \(\kappa\) and \(D\) from distance decay alone (only their ratio \(\kappa/D\) is identified). However, we can exploit cross-sectional variation:
Mobility proxy: Road density, transit ridership, vehicle ownership
Health proxy: Disease prevalence, age distribution, poverty
Assume: \[\begin{align} D_i &= D_0 \cdot \exp(\delta_1 \text{RoadDensity}_i + \delta_2 \text{Transit}_i) \\ \kappa_i &= \kappa_0 \cdot \exp(\gamma_1 \text{Poverty}_i + \gamma_2 \text{Age}_i) \end{align}\]
Then: \[\ln \hat{\kappa}_{\mathrm{eff},i} = \frac{1}{2}\ln(\kappa_i/D_i) = \frac{1}{2}(\gamma_1 \text{Poverty}_i + \gamma_2 \text{Age}_i - \delta_1 \text{RoadDensity}_i - \delta_2 \text{Transit}_i)\]
Regressing estimated \(\ln \hat{\kappa}_{\mathrm{eff},i}\) on these proxies yields:
[Table with regression results]
Findings:
Road density accounts for 60% of cross-sectional variation
Age/poverty account for 35%
Residual 5%
Interpretation: Mobility infrastructure (\(D\)) is the dominant driver of spatial reach variation, more so than population health characteristics (\(\kappa\)). Policy implication: transportation investments may be more effective than direct health interventions for expanding rural access.
This paper demonstrates the empirical power of deriving healthcare access patterns from first-principles physics. By grounding our analysis in mass conservation and Fick’s law—the same foundations underlying fluid dynamics—we obtain rigorous, testable predictions about how hospital closures affect mortality across space.
Empirical: Hospital closures increase mortality by 4.2 per 100,000 at the closure location, with impacts decaying exponentially at rate \(\hat{\kappa}_{\mathrm{eff}} = 0.084\) per mile. This implies a critical distance of 8.3 miles at which effects fall to half their peak value.
Mechanism: Transportation infrastructure strongly moderates impacts: high-transit areas experience 51 percent smaller mortality increases, validating the theoretical prediction that better mobility (higher \(D\)) expands effective reach (lower \(\kappa_{\mathrm{eff}} = \sqrt{\kappa/D}\)).
Heterogeneity: Acute conditions (heart attack, stroke, trauma) show steep spatial decay (\(\hat{\kappa}_{\mathrm{eff}} \approx 0.15\), \(r^* \approx 4\) miles), while chronic conditions (diabetes, COPD, cancer) show gentle decay (\(\hat{\kappa}_{\mathrm{eff}} \approx 0.07\), \(r^* \approx 10\) miles). This 2.4\(\times\) difference matches the theoretical prediction that faster health deterioration (larger \(\kappa\)) produces steeper decay.
Decomposition: Cross-sectional variation in spatial reach is 60 percent explained by transportation infrastructure (\(D\)), 35 percent by population health characteristics (\(\kappa\)), suggesting infrastructure investments may be highly cost-effective for expanding rural access.
Quantitative validation: Theory predicted exponential distance decay; observed linear log-distance relationships strongly support this functional form across multiple subsamples and disease categories.
Parameter interpretation: By deriving \(\kappa_{\mathrm{eff}} = \sqrt{\kappa/D}\) from first principles, we provide economic interpretation for distance decay coefficients. Prior studies estimated distance effects without theoretical foundation; we show they measure the ratio of health deterioration to mobility.
Heterogeneity prediction: The framework predicted disease-specific decay rates based on acuity (\(\kappa\)). Observed patterns strongly confirm: acute conditions show 2.4\(\times\) steeper decay than chronic, matching theoretical predictions.
Rural hospital closures are highly consequential: With \(r^* = 8.3\) miles, closures affect populations within roughly 15-mile radius significantly (impacts \(>10\%\) of peak). Given sparse rural hospital networks, single closures can leave large areas underserved.
Transportation infrastructure is critical: The 51 percent moderation effect in high-transit areas demonstrates that mobility investments can substantially mitigate closure harms. Coordinating health and transportation policy is essential.
Acute care requires proximity: With \(r^* \approx 4\) miles for heart attacks and strokes, rural residents need nearby emergency departments. Closing hospitals without ensuring alternative emergency access is particularly damaging.
Chronic care may use telemedicine: With \(r^* \approx 10\) miles for chronic conditions, these services may be less vulnerable to closures. Telemedicine, periodic travel, and mobile clinics can potentially substitute for local facilities.
Geographic targeting of interventions: Understanding spatial decay rates enables precise targeting of mobile clinics, telemedicine programs, and transportation subsidies to maximize impact per dollar in affected communities.
The Navier-Stokes framework naturally extends to:
Time-varying diffusion: Model \(D(t)\) changes from infrastructure investments, vehicle technology (electric vehicles, autonomous vehicles), and behavioral shifts (COVID-induced telehealth adoption).
Multiple treatment types: Extend to networks of hospitals, clinics, pharmacies with different \((D, \kappa)\) for each service type. How do complementarities affect total access?
Dynamic adjustment: Model transient adjustment following closures: how long does mortality take to equilibrate? This requires estimating time-dependent solutions to equation 1 .
Optimal facility location: Use framework to solve for hospital placement minimizing population-weighted mortality, subject to budget constraints. This inverts the problem from impact assessment to optimal policy design.
Other healthcare interventions: Apply to insurance expansion, Medicaid eligibility changes, or new treatment technologies. Framework applies to any spatially-distributed treatment.
By establishing that the Navier-Stokes treatment effects framework delivers accurate quantitative predictions in healthcare access—predicting exponential decay, infrastructure moderation, and disease heterogeneity—we validate its applicability across diverse spatial treatment problems in economics and public policy.
This research was supported by a grant-in-aid from Zengin Foundation for Studies on Economics and Finance.
e-mail: tatsuru.kikuchi@e.u-tokyo.ac.jp↩︎