Next Article in Journal
Evaluation of Atmospheric Environmental Regulations: The Case of Thermoelectric Power Plants
Previous Article in Journal
Modeling and Assessment of PM10 and Atmospheric Metal Pollution in Kayseri Province, Turkey
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Initial Conditions and Resilience in the Atmospheric Boundary Layer of an Urban Basin

1
Departamento de Física, Facultad de Ciencias Naturales, Matemáticas y Medio Ambiente, Universidad Tecnológica Metropolitana, Las Palmeras 3360, Ñuñoa, Santiago 7750000, Chile
2
Departamento de Industria, Facultad de Ingeniería, Universidad Tecnológica Metropolitana, Santiago 7750000, Chile
*
Author to whom correspondence should be addressed.
Atmosphere 2023, 14(2), 357; https://doi.org/10.3390/atmos14020357
Submission received: 27 December 2022 / Revised: 6 February 2023 / Accepted: 6 February 2023 / Published: 10 February 2023
(This article belongs to the Special Issue Comprehensive Modeling of Air Pollution and Its Weather Effects)

Abstract

:
The possibilities of micrometeorological resilience in urban contexts immersed in a basin geographical configuration are investigated. For this purpose, time series data with measurements of meteorological variables (temperature, magnitude of wind speed and relative humidity) and atmospheric pollutants (PM2.5, PM10, CO) are analyzed through chaos theory, calculating the coefficient of Lyapunov (λ), the correlation dimension (Dc), the Hurst coefficient (H), the correlation entropy (SK), the fractal dimension (D) and the Lempel–Ziv complexity (LZ). Indicators are built for each measurement period (2010–2013 and 2017–2020), for each locality studied and located at different heights. These indicators, which correspond to the quotient between the entropy resulting from the meteorological variables and that of the pollutants, show sensitivity to height. Another important indicator, for identical measurement conditions, arises from the calculation of the fractal dimensions of the meteorological variables and that of the pollutants, which allows for comparative studies between the two periods. These indicators are conclusive in pointing out that, in a large city with basin geographical characteristics, subjected to an intensive urbanization process, there is no micrometeorological resilience and a great variation occurs in the initial conditions.

1. Introduction

1.1. Resilience

Human interaction with nature has always had consequences [1]. The difference with today’s times is that generally, in the past, nature could reduce or absorb these effects in relatively short times. Various factors have strained this capacity to the point that the disturbances, which now occur continuously and in short periods of time, are not assimilated in the short term.
Resilience is the term used in the ecology of communities and ecosystems to indicate their ability to absorb disturbances, keeping their structural, dynamic and functional characteristics practically intact; being able to return to the situation prior to the disturbance after the cessation of it [2]. Thus, resilience is a measure of the magnitude of disturbances that a system can absorb to move from one equilibrium to another.
The idea of resilience can arise from a reductionist approach; the system is limited to the sum of the interactions between its parts. However, when a complex system manifests emergence, that is, a general behavior that cannot be described by the sum of the interactions of its components, a systems-based approach can be more fruitful. It is observed in nature, for example, that the tendency for the mixed phase of arctic clouds to persist despite their inherent microphysical instability suggests emergent qualities. It also evokes self-organizing properties, defined as “internal, local interactions of processes that give rise to global order”. The nature of these interactions results in a robust system that is resistant to disturbance.
Self-organization prevails in a variety of natural and man-made factors of systems. Examples include oscillating chemical reactions, behavior among bird populations, predator–prey interactions in the fields of ecology and cloud physics [3], light amplification by stimulated emission of radiation, and computer network theory. In the subtropics and mid-latitudes, cloud fields associated with mesoscale convection often occur as closed or open cellular patterns, exhibiting system-wide ordering that selects the preferred state based on, among other things: environmental conditions or due to external forces. Small perturbations often strengthen the robustness of the state by allowing further exploration of phase space in the proximity of the attractor, while large perturbations can cause the system to transition from one preferred state to another [4].

1.2. Cities and Risk

Risk, threat, vulnerability, and disaster are terms from the resilience literature that are related to one another. For example, risk can be defined in terms of hazards and vulnerability. The probability of a dangerous event occurring only becomes a risk if it can negatively affect people, communities or systems, that is, if these actors are susceptible to the impacts of threats (vulnerability). Disaster risk, for its part, refers to the possible negative effects of a hazard, which are determined in relation to risk exposure and the ability to act on it [5].
Urban resilience can be understood as the permanent capacity of cities to absorb, adapt, transform and prepare for shocks and stands out in the economic, social, institutional and environmental dimensions, with the objective of maintaining the functions of a city and improving the response to future crises [5].
From another perspective, it is argued that economic activities are sustainable only if the ecosystems that support life, and on which they are dependent, have an adequate level of resilience. Approaching the problem from a multivariate perspective is an arduous task. At present we have been learning that the environmental dimension is decisive for the viability of a city [6]. Within this dimension, this study will limit the problem to urban centers: the meteorology of the limit layer of the atmosphere and the disturbance due to air pollution by particulate matter (PM2.5 and PM10) and carbon monoxide (CO). Many questions arise from the previous considerations: is it possible to build indicators of disturbance of the atmosphere? What kind of processes can be induced? Can a system return to its initial condition if the disturbance is increasing over time and incorporates new elements such as buildings, change in surface roughness, decay of humidity, etc.?
Through the chaos theory, applied to time series data of urban micrometeorology [7,8,9,10,11,12,13] and of atmospheric pollutants, is it possible to obtain information on the resilient behavior of the atmospheric boundary layer, comparing different periods?

1.3. Entropy

As noted, all systems present in nature are subjected to irreversible processes [14,15,16]. This occurs when a system and its surroundings cannot return to their initial state [17]. It is observed that the phenomenon of irreversibility according to Prigogine [17] has a constructive character, highlighting the “creative role of time”. Therefore, it supposes, at least at the macroscopic level, a kind of anti-entropy: the non-equilibrium universe is a connected universe. This reveals to us that in the phenomena of nature there is connectivity and irreversibility. A magnitude that accounts for systems that have these characteristics is entropy. This is a fundamental variable that has different definitions and is measured in different ways, depending on the different fields in which it is applied [16,18,19].
From a historical perspective, in classical physics the entropy of a physical system is proportional to the amount of energy not available to do physical work. In quantum mechanics, von Neumann extended the notion of entropy to quantum systems by means of the density matrix [20]. In probability theory, the entropy of a random variable measures the uncertainty regarding the value that should be assumed by the variable [21]. In information theory, the compression entropy of a message (a computer file) quantifies the amount of information that is carried by the message in terms of the best lossless compression ratio [22]. In dynamic systems theory, entropy quantifies the exponential complexity of systems or the average flow of information per unit of time [23]. In sociology, entropy is the natural decay of the structures (such as laws, organization, and conventions) of a social system. Additionally, in the usual sense, entropy means disorder or chaos. By saying that entropy is a measure of disorder, it is implied that the highest entropy is the greatest disorder, which appears when it is considered as a measure of the unavailability of a system’s energy to do work. It is a parameter that represents the state of disorder of an atomic, ionic, or molecular system. It is also a measure of disorder in the universe or simply a measure of disorder in any system [24,25].

1.4. Kolmogorov Entropy and Its Relation to Information Loss

Following what has been indicated by Farmer, one of the basic differences between chaotic and predictable behavior is that chaotic trajectories continuously generate new information while predictable trajectories do not [26,27]. Metric entropy makes this notion more rigorous. In addition to providing a good definition of chaos, metric entropy provides a quantitative way to describe how chaotic a dynamical system is. In the Kolmogorov entropy [28,29], SK is the average loss of information [22] when “l” (side of the cell in units of information) and τ (time) become infinitesimal [30]:
S K = lim τ 0   lim l 0 lim n n τ 1 i = 1 n P i logP i
SK has units of information bits per second and bits per iteration in the case of a discrete system [31,32]. The limit process of Equation (1) develops according to the order: (i) n → ∞, (ii) l → 0, canceling the dependency of the selected partition (n is the number of cells or partitions) and (iii) τ → 0, for continuous systems. The Kolmogorov entropy difference, ΔSK = SKn+1 − SKn between two neighboring cells, gives the required complementary information about the cell (in+1) in which the system will be in the future. The difference gives the loss of information, in time, of the system [30].
In summary, for the calculation of the Kolmogorov entropy, it is first verified that the entropy is between zero and infinity (0 < SK < ∞), which allows verifying the presence of chaotic behavior. If the Kolmogorov entropy is equal to 0, no information is lost and the system is regular and predictable. If SK     ∞, the system is completely random, and it is impossible to make any predictions. Second, the amount of information required to predict the future behavior of two interacting systems is determined, in this case, the atmosphere and the hourly concentration of pollutants (PM10, PM2.5 and CO). In this way, the rate at which the system loses (or outdates) information over time can be estimated. Finally, the horizon of maximum predictability of the system can be established. This horizon is a limit frontier from which it is not possible to make predictions or formulate new scenarios [29]. The loss of information can be determined according to the equation:
I = I NEW I OLD = λ i 0 t log 2
λ is the Lyapunov exponent, <ΔI> in bits/h, is the loss information. Two types of <ΔI> can be calculated: one for the contribution of pollutants, and another for the sum of the loss of information of each meteorological variable (MV): temperature (T), magnitude of wind speed (WS) and relative humidity (HR) [30].

2. Materials and Methods

2.1. Materials

In Chile, there is an online public network for monitoring meteorological and pollutant variables [33] that has extensive periods of data records that form time series. The field recording of data serves as a starting line for this research. To analyze the data, we followed the line of chaos theory applied to time series [27,28,29,30,31,32,33,34,35,36,37,38,39,40,41]. Many authors have approached the treatment of time series in this way [23,42,43,44,45,46,47,48,49,50,51]. From the extensive data records, measurements of meteorological variables (magnitude of wind speed, relative humidity, temperature) and air pollutants (PM2.5 particulate matter, PM10 and carbon monoxide CO) constituted in time series for different periods were selected in the first approximation [51,52,53,54,55]. The analysis of these series, through a software that explores their chaoticity, allows to determine if their nature is chaotic from the ranges of five parameters characteristic of chaos theory: Lyapunov’s coefficient (λ > 0), Hurst’s Coefficient (0.5 < H < 1), correlation dimension (Dc < 5), correlation entropy (SK > 0), and complexity Lempel–Ziv (LZ > 0). Finally, the comparison between the entropies resulting from the meteorological variables and that of the pollutants gives an indicator of variation by morphological geography.

2.2. Methods

The flowchart, Figure 1, indicates the stages of the procedure to be followed once the time series of the pollutants and the meteorological variables have been obtained [30].

2.3. Study Area

The city of Santiago is located at 33.5° S and 70.8° W. It contains a population of 7,508,334 inhabitants, which represents 40% of the total population of the country, on a surface of approximately 641 km2. It is located in the middle of the country, at a height of about 520 m.a.s.l. The altitude above sea level increases from west to east. It is surrounded by two mountain chains: The Andes and the Coastal Mountain range. Its climate is Mediterranean. The driest and warmest months are from December to February, reaching maximum temperatures of about 35 °C in the shade (air temperature in the sun), Figure 2. Given its topography and the dominant meteorological conditions, there is in general a strong horizontal and vertical dispersion of pollutants generated by an important number of sources in the city (heating, vehicles, industries, etc.), especially during autumn (20 March–21 June) and winter (21 June–23 September) [51,56,57,58,59,60]. The emissions have a tendency to increase given the increasing population density, which implies an increase in fixed and mobile sources. In addition, the number of vehicles has increased rapidly in recent years.

3. Results

Two tables (Table 1 and Table 2) were built, one for the entropies of the meteorological variables and another for the entropies of the pollutants. Data were extracted from references [49] (page 175) and [51](page 11). The heights, with respect to sea level, are the same for the two periods (2013–2013 and 2017–2020) of the study and correspond to that of the monitoring stations.

3.1. Meteorological Variables

Figure 3 presents the distribution of entropies according to heights and periods of measurement of the meteorological variables.

3.2. Pollutants

Figure 4 presents the distribution of entropies according to heights and periods of the pollutants.

3.3. Relationship between SK,MV and SK,P

It can be establish a relationship between CK= SK,MV/SK,P = S K . MV , i / S K . P , i COMMUNES and the height of the different monitoring stations of meteorological and pollutant variables and, in addition, of the measurement periods, what appears in Table 3 and is plotted in Figure 5.
Figure 5 represents the relationship between CK and the height:

4. Discussion

A scalar time series is one-dimensional. There are important aspects of the scale of measurement and they are related to the expression of the size and to the topological dimension: the value of the magnitude depends on the value of the scale, but the magnitude of the topological dimension is independent of the scale (for example, length, area and volume will appear more fractional due to scale changes).
As fractals are made up of smaller and smaller elements of themselves, a set of recursive or self-similar points at any scale is arrived at. This set determines a geometry that is contained in a certain dimension. Mandelbrot, points out that a fractal is, by definition, a set whose Hausdorff–Besicovitch dimension is strictly greater than its topological dimension. Thus, the key mathematical property of an essentially fractal object is that its fractal metric dimension is a rational number greater than its topological dimension. The dimension of the time series graph (the trace) is the fractal dimension of the graph generated by the time series. If the series is a scalar observable, then this dimension will be between 1.0 for a very smooth curve and 2.0 for a very erratic one [23]. As the fractal dimension increases, the geometry of the fractal object contained in the time series is more complex, the described process is more complex or more chaotic. Considering that D = 2 − H, where D is the fractal dimension and H is the Hurst exponent; due to the values obtained, the fractal dimension of the time series for the period 2010–2013 are closer to the line and those for the period 2017–2020 are further from the line (Table A1, Appendix).
Each of the time series studied had a large and positive Lyapunov exponent, which means that each is chaotic and will usually have a strange attractor with non-integer dimension, as shown by the calculations. In the case of the time-delay embedding space, it is a reconstructed state space chosen with the minimum dimension for which important dynamical and topological properties are preserved.
If the Hurst coefficient is between 0.5 < H < 1, the time series are persistent, that is, characterized by long-term memory effects. Theoretically, what happens at a given moment affects the future. There is no characteristic time scale, which is the key property of fractal time series. In terms of chaotic dynamics there is sensitivity about the initial conditions. The persistent series are the most common found in nature.
The calculations presented in Table A1 [30,51] (in the Appendix) show that the Hurst exponent and the Lempel–Ziv complexity of the wind speed magnitude is, in general, higher in the period 2017–2020 compared to the period 2010–2013. The wind shows a more turbulent behavior for the period 2017–2020.
According to the Kolmogorov cascade effect, the fully developed turbulent medium is characterized by two quantities, the average rate of energy dissipation, εd, and the kinematic viscosity ν. The dimensionality of εd is energy/time/mass, or L 2 T 3 , and the dimensionality of ν is L 2 T 1 . We can combine these two quantities to find the length scale:
l k = ν 3 ε d 1 / 4
known as the Kolmogorov or dissipational length scale. The length scale is the size of the smallest eddies in the fluid. Eddies smaller than the Kolmogorov scale rapidly dissipate their kinetic energy by viscous heat and disappear [28]. The eddies of size lK rotate with a speed:
u K = ν ε d 1 / 4
and dissipates its energy in a time, roughly, equal to
τ K = ν ε d 1 / 2
In Earth’s atmosphere, ε d = 10 c m 2 s 3 and ν   ~   0.1 c m 2 s . Thus, the smallest eddy in the Earth’s atmosphere is l K   ~   0.1   c m , which is significantly larger than the mean free path λ ~   10 4 cm . Smaller eddies rotate around the surroundings with speed u K ~ 1 c m s , dying according to a decay time of   τ K ~ 0.1   s .
The Kolmogorov entropy of the time series of this study has units S K bits h . When transforming their units by means of Landauer’s principle,   S K J K   h remains, which is equivalent to energy/(temperature time), dimensionally:
M T L 2 t 3 ~ M T * average   power   dissipation   rate
In general, the rate of change of C K = S K , MV / S K , P with height is:
d dh C K = d dh S K , MV S K , P = S K , P dS K , MV dh S K , MV dS K , P dh S K , P 2 = dS K , MV dh S K , MV S K , P dS K , P dh S K , P
Series 1: 450 m < h < 550 m, very close to the ground, for the period 2010–2013:
d dh S K , MV S K , P = dS K , MV dh S K , MV S K , P dS K , P dh S K , P < 0
Figure 5 exhibits, in a first approximation, a strong variation of CK in certain ranges of height, which could be compatible with the generation, by increases in building height in the period 2017–2020 (Series 2 furthest from equilibrium ( S K , M V S K , P ~ 1 ) between the entropies of the selected pollutants and the meteorological variables), of a more turbulent wind regime in the basin.
Since SK,P is always greater than zero, inequality (7) is satisfied if:
dS K , MV dh < S K , MV S K , P dS K , P dh
with:
S K , MV S K , P > 0
1 S K , MV dS K , MV dh < 1 S K , P dS K , P dh
      d   ln   S K , MV dh < d   lnS K , P dh
Very close to the ground, the variation with the height of CK is negative and of very large magnitude, it decreases towards a relative minimum. Urban weather is more stable or “strong”, which makes it less vulnerable to the effect of air pollutants. As an initial condition, this “strength” has a detrimental effect on air pollutants. As height increases, this variation smooths out. The variation with the height of the ln of the entropy of the meteorological variables is less than the variation with the height of the ln of the entropy of the pollutants, that is, the meteorology of the period 2010–2013 is dynamically more persistent and balances the polluting system generated through the term σ (with units of length−1):
d   ln   S K , MV dh + σ = d   lnS K , P dh
It is possible to reduce the system, in the form:
d ln S K , MV + σ dh = d   lnS K , P d ln S K , MV d   lnS K , P = σ dh
d   ln S K , MV S K , P = σ dh S K , MV S K , P = e σ h S K , MV = S K , P e σ h
Verification, differentiating with respect to h:
dS K , MV dh = dS K , P dh e σ h S K , P σ e σ h
dS K , MV dh = dS K , P dh S K , MV S K , P S K , P σ S K , MV S K , P
dS K , MV dh = dS K , P dh S K , MV S K , P S K , MV σ
1 S K , MV dS K , MV dh = 1 S K , P dS K , P dh σ d   ln   S K , MV dh + σ = d   lnS K , P dh
Series 2: 450 m < h < 550 m, very close to the ground, for the period 2017–2020:
d dh S K , MV S K , P = dS K , MV dh S K , MV S K , P dS K , P dh S K , P > 0
from where:
dS K , MV dh > S K , MV S K , P dS K , P dh   with   S K , MV S K , P > 0
1 S K , MV dS K , MV dh > 1 S K , P dS K , P dh d   ln   S K , MV dh > d   lnS K , P dh
In this case, it follows:
d   ln   S K , M V d h = δ + d   ln S K , P d h
Very close to the ground, the variation with height of CK is positive and of very large magnitude and grows towards a relative maximum. Urban weather is more variable, turbulent or less “strong”, making it more vulnerable to the effect of more stable air pollutants. As height increases, this variation smooths out. The variation with the height of the ln of the entropy of the meteorological variables is greater than the variation with the height of the ln of the entropy of the pollutants, that is, the pollutants are dynamically more persistent, balancing the system with the term, according to height, δ (with units of length−1).
Some turbulence breaks out around 700 m, more moderate than between 450 and 500 m, and at higher altitudes the processes represented by both series tend to stabilize at different CK’s depending on the study periods. It is possible to model, in a very approximate way, according to Equations (19) and (20) the processes described by CK in each period:
2010–2013 (Series 1):
C K h = S K , MV S K , P = 0.8 + 3.5 * e 6 h 0.2 * sin 3.5 * 2 * h 0.2 / 6
2017–2020 (Series 2):
C K h = S K , MV S K , P = 0.75 1.5 * e 6 h 0.2 * sin 3.5 * 2.2 * h 0.2
obtaining Figure 6.
Figure 7 is the graphical representation of Equations (19) and (20). It allows to visualize the effect on CK of the heights close to the Earth’s surface where the density of the atmosphere is higher with higher kinematic viscosity. The opposite happens when the height is greater, although the volume is also greater, the kinematic density is lower, so the energy exchanges are smoother and tend to balance compared to the atmosphere closer to the ground.
Figure 7 conforms to a phenomenology of the type described by the Kolmogorov cascade, as represented in Figure 8.
In addition, the analysis of the ratio between the entropies of the urban meteorology variables and those of the pollutants shows sensitivity according to the measurement periods. If the system is considered in the period 2010–2013, as shown in the representation in Figure 5 of the ratio with the height, it is the initial condition of the total period of the study. The variation of this initial condition, due to systematic anthropic disturbance over time, as shown in Figure 5 of the 2017–2020 period, produces a different behavior of the system. Variations of the initial condition, induced by the increasing contamination of the atmosphere, determine different evolutions of the system with a “relatively asymptotic” behavior of CK as the height increases. It stabilizes at higher altitudes by increasing the volume of the atmosphere and the proximity to the upper atmospheric layer (which has a different thermodynamics).
Although urban meteorology has its natural modes of relationship (with solar radiation, volcanoes, geography, seas, etc.) and entropy variations, with respect to the historical average, the interactions were naturally processed in ancient and resilient periods. Human activity is a disruption and contributes to the imbalance of the system [30,57,58,59,60,61,62,63,64,65,66].

5. Conclusions

The data from the two time periods, 2010–2013 and 2017–2020, from urban meteorology measurements (three time series) and pollutants (three time series) correspond to a continuous and hourly instrumental record. The analysis applied to the set of variables that the system represents proves that it is a complex system, of non-linear dynamics, very sensitive to variations in the initial conditions. Small variations in these initial conditions can imply large differences in future behavior, making long-term prediction impossible. The variation with the height of CK indicates that it varies strongly and positively (dCK/dh) at low altitudes in the period 2017–2020 (intensification of high-rise buildings), unlike the period 2010–2013 (dCK/dh < 0). The process declines asymptotically with height, with a gap between periods, but the 2017–2020 period is the one with the greatest drop. According to Table A2 of the Appendix, in the period 2017–2020 there is a greater complexity and persistence of the wind (greater turbulence). The entropy, the Hurst exponent, the fractal dimension, and the loss of information indicates the impossibility of returning the studied system (corresponding to a measurement context and a geography) to its initial condition because there are elements of perturbation, continuous action, and that are the result of human activity. From the perspective of the ground measurements and the general mathematical model of analysis, a return to the initial condition is not concluded, which is one of the basic characteristics of a resilient natural system, in this case referring to the period of this study of the atmosphere. The idea of a resilient city, as an end in itself, makes no sense as it is inserted in its interaction and connectivity with nature.

Author Contributions

Conceptualization, P.P. and E.M.; methodology, P.P. and E.M.; software, P.P.; validation, P.P. and E.M.; formal analysis, P.P. and E.M.; investigation, P.P. and E.M.; resources, E.M., V.F. and C.P.; data curation, E.M.; writing—original draft preparation, P.P. and E.M.; writing—review and editing, P.P. and E.M.; visualization, E.M., V.F. and C.P.; supervision, P.P.; project administration, P.P.; funding acquisition, P.P. All authors have read and agreed to the published version of the manuscript.

Funding

Project supported by the Competition for Research Regular Projects from the year 2020, code LPR20-02, Universidad Tecnológica Metropolitana.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

The data were obtained from the public network for online monitoring of air pollutant concentration and meteorological variables. In Chile: The network is distributed throughout all of Chile, without access restrictions. It is the responsibility of SINCA, the National Air Quality Information System, dependent on the Environment Ministry of Chile [33]. The data for the two study periods will be available for free-use on the WEB page: https://sinca.mma.gob.cl (accessed on 23 April 2022).

Acknowledgments

To the Research Directorate of the Universidad Tecnológica Metropolitana that made possible the progress of this research.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. All the values of the Hurst exponent (H), the fractal dimension (D) for all the variables of interest by studied community and the two periods (2010–2013 and 2017–2020).
Table A1. All the values of the Hurst exponent (H), the fractal dimension (D) for all the variables of interest by studied community and the two periods (2010–2013 and 2017–2020).
PM10PM2.5COTHRWV
EML2010–2013
H0.9670.9730.9590.9890.9910.976
D1.0331.0271.0411.0111.0091.024
2017–2020
H0.9220.9630.9330.9150.9420.975
D1.0781.0371.0671.0851.0581.025
EMM2010–2013
H0.9720.9770.9810.9910.9900.980
D1.0281.0231.0191.0091.0101.02
2017–2020
H0.9060.9830.9330.9170.9410.976
D1.0941.0171.0671.0831.0591.024
EMN2010–2013
H0.9720.9740.9530.9890.9910.968
D1.0281.0261.0471.0111.0091.032
2017–2020
H0.9290.9600.9330.9160.9420.973
D1.0711.041.0671.0841.0581.027
EMO2010–2013
H0.9650.9550.9370.9920.9890.968
D1.0351.0451.0631.0081.0111.032
2017–2020
H0.9360.9250.9330.9190.9420.974
D1.0641.0751.0671.0811.0581.026
EMS2010–2013
H0.9690.9730.9530.9900.9920.957
D1.0311.0271.0471.0101.0081.043
2017–2020
H0.9210.9750.9330.9150.9420.976
D1.0791.0251.0671.0851.0581.024
EMV2010–2013
H0.9670.9700.9520.9890.9890.956
D1.0331.031.0481.0111.0111.044
2017–2020
H0.9310.9660.9330.9190.9420.975
D1.0691.0341.0671.0811.0581.025
Table A2. Comparative table of the amount of wind speed according to persistence, H, and LZ complexity, where > means an increase, and = means the same value.
Table A2. Comparative table of the amount of wind speed according to persistence, H, and LZ complexity, where > means an increase, and = means the same value.
StationsEMLEMMEMVEMNEMSEMO
PeriodsH; LZH; LZH; LZH; LZH; LZH; LZ
2010–20130.976; 0.3200.980; 0.5580.956; 0.3250.968; 0.2860.957; 0.2930.968; 0.538
2017–20200.975 (=); 0.551 (>)0.976 (=); 0.557 (=)0.975 (>); 0.544 (>)0.973 (>);0.539 (>)0.976 (>); 0.556 (>)0.974 (>); 0.537 (=)

References

  1. Cox, D.T.C.; Gaston, K.J. Human–nature interactions and the consequences and drivers of provisioning wildlife. Phil. Trans. R. Soc. B 2018, 373, 20170092. [Google Scholar] [CrossRef] [PubMed]
  2. Holling, C.S. Resilience and Stability of Ecological Systems. Annu. Rev. Ecol. Syst. 1973, 4, 1–23. [Google Scholar] [CrossRef]
  3. Morrison, H.; Gijs de Boer, G.; Feingold, G.; Feingold, G.; Harrington, J.; Shupe, M.D.; Sulia, K. Resilience of persistent Arctic mixed-phase clouds. Nat. Geosci. 2012, 5, 11–17. [Google Scholar] [CrossRef]
  4. González, J.E.; Ramamurthy, P.; Bornstein, R.D.; Chen, F.; Bou-Zeid, E.R.; Ghandehari, M.; Luvall, J.; Mitra, C.; Niyogi, D. Urban climate and resiliency: A synthesis report of state of the art and future research directions. Urban Clim. 2021, 38, 100858. [Google Scholar] [CrossRef]
  5. Figueiredo, L.; Honiden, T.; Schumann, A. Indicators for Resilient Cities. In OECD Regional Development Working Papers; No. 2018/02; OECD Publishing: Paris, France, 2018. [Google Scholar] [CrossRef]
  6. Travaglini, S.; Noble, J.; Ross, P.G.; Dharan, C.K.H. Mycology Matrix Composites. In Proceedings of the American Society for Composites—Twenty-Eighth Technical Conference (Annual Technical Conference 28th, American Society for Composites), State College, PA, USA, 9–11 September 2013; Curran Associates, Inc., Red Hook: New York, NY, USA, 2013; pp. 517–535, ISBN 9781629931432. (British Library Conference Proceedings). [Google Scholar]
  7. Mauree, D.; Coccolo, S.; Deschamps, L.; Loesch, P.; Becquelin, P.; Scartezzini, J.L. Mobile Urban Micrometeorological Monitoring (MUMiM). J. Phys. Conf. Ser. 2019, 1343, 012014. [Google Scholar] [CrossRef]
  8. Foken, T. Micrometeorology; Springer: Berlin/Heidelberg, Germany, 2008. [Google Scholar]
  9. Stull, R.B. Introduction to Boundary Layer Meteorology; Kluwer Academic: Alphen aan den Rijn, The Netherlands, 1988; 666p. [Google Scholar]
  10. Klausner, Z.; Ben-Efraim, M.; Arav, Y.; Tas, E.; Fattal, E. The Micrometeorology of the Haifa Bay Area and Mount Carmel during the summer. Atmosphere 2021, 12, 354. [Google Scholar] [CrossRef]
  11. Garratt, J.R.; Pearman, G.I. Retrospective Analysis of Micrometeorological Observations Above an Australian Wheat Crop. Bound.-Layer Meteorol. 2020, 177, 613–641. [Google Scholar] [CrossRef]
  12. Landsberg, H.E. Micrometeorological temperature differentiation through urbanization. In Urban Ctimates, Technical Note, No. 108; World Meteorologicai Organization: Geneva, Switzerland, 1970; pp. 129–136. [Google Scholar]
  13. Landsberg, H.E.; Maisel, T.N. Micrometeorological observations in an area of urban growth. Bound.-Layer Meteorol. 1972, 2, 365–370. [Google Scholar] [CrossRef]
  14. Bénard, H. Les tourbillons cellulaires dans une nappe liquide propageant de la chaleur par convection en régime permanent. Ann. Chim. Phys. 1901, 23, 62–164. [Google Scholar] [CrossRef]
  15. Brillouin, L. Science and Information Theory, 2nd ed.; Academic Press: New York, NY, USA, 1962; 304p. [Google Scholar]
  16. Haddad, W. Thermodynamics: The Unique Universal Science. Entropy 2017, 19, 621. [Google Scholar] [CrossRef]
  17. Prigogine, I. Introduction to Thermodynamics of Irreversible Processes; Thomas, C.C., Ed.; ACS Publications: Springfield, IL, USA, 1955; 115p. [Google Scholar]
  18. Sinai, Y.G. On the concept of entropy of a dynamical system. Dokl. Akad. Nauk SSSR 1959, 124, 768–777. [Google Scholar]
  19. Dewar, R.C. Maximum entropy production and the fluctuation theorem. J. Phys. A Math. Gen. 2005, 38, 1371–1381. [Google Scholar] [CrossRef]
  20. Bengtsson, I.; Życzkowski, K. Geometry of Quantum States: An Introduction to Quantum Entanglement, 301, 2nd ed.; Cambridge University Press: Cambridge, UK, 2017. [Google Scholar]
  21. Parker, D. The H-Theorem, Molecular Disorder and Probability: Perspectives from Boltzmann’s Lectures on Gas Theory PhilSci Arch 2009, OAI–Open Archives Initiative. Available online: http://philsci-archive.pitt.edu/id/eprint/5005 (accessed on 26 November 2020).
  22. Shannon, C. A mathematical theory of communication. Bell Syst. Tech. 1948, 27, 379–423. [Google Scholar] [CrossRef]
  23. Sprott, J.C. Chaos and Time-Series Analysis; Oxford University Press: Oxford, UK, 2003; 528p. [Google Scholar]
  24. Ruelle, D.P. Extending the definition of entropy to nonequilibrium steady states. Proc. Natl. Acad. Sci. USA 2003, 100, 3054–3058. [Google Scholar] [CrossRef]
  25. Lin, Y.; Lin, C.; Wang, D. Understanding Atmospheric Behaviour in Terms of Entropy: A Review of Applications of the Second Law of Thermodynamics to Meteorology. Entropy 2011, 13, 211–240. [Google Scholar] [CrossRef]
  26. Farmer, J.D. Chaotic attractors of an infinite dimensional dynamical system. Physica D 1982, 4, 366–393. [Google Scholar] [CrossRef]
  27. Farmer, J.D.; Otto, E.; Yorke, J.A. The dimension of chaotic attractors. Physica D 1983, 7, 153–180. [Google Scholar] [CrossRef]
  28. Kolmogorov, A.N. On Entropy per unit Time as a Metric Invariant of Automorphisms. Dokl. Akad. Nauk SSSR 1959, 124, 754–755. [Google Scholar]
  29. Martínez, J.A.; Vinagre, F.A. La Entropía de Kolmogorov; su Sentido Físico y su Aplicación al Estudio de Lechos Fluidizados 2D. Departamento de Química Analítica e Ingeniería Química, Universidad de Alcalá, Alcalá de Henares, Madrid, España. Available online: https://www.academia.edu/247937219/07/2019 (accessed on 20 September 2020).
  30. Pacheco, P.; Mera, E. Relations between Urban Entropies, Geographical Configurations, Habitability and Sustainability. Atmosphere 2022, 13, 1639. [Google Scholar] [CrossRef]
  31. Shaw, R. Strange attractors, chaotic behavior and information flow. Z. Naturforsch. A 1981, 36, 80–112. [Google Scholar] [CrossRef]
  32. Cohen, A.; Procaccia, I. Computing the Kolmogorov entropy from time signals of dissipative and conservative dynamical systems. Phys. Rev. A Gen. Phys. 1985, 31, 1872–1882. [Google Scholar] [CrossRef] [PubMed]
  33. MMA, Sistema de Información Nacional de Calidad del Aire (Ministerio del Medioambiente de Chile). Available online: https://sinca.mma.gob.cl/index.php (accessed on 23 April 2022).
  34. Takens, F. Detecting Strange Attractors in Turbulence. In Dynamical Systems and Turbulence, Warwick 1980; Rand, D., Young, L.S., Eds.; Springer: Berlin, Germany, 1981; pp. 366–381. [Google Scholar] [CrossRef]
  35. Eckmann, J.P.; Oliffson, S.; Kamphorst, S.; Ruelle, D.; Ciliberto, C. Lyapunov exponents from time series. Phys. Rev. A 1986, 34, 4971–4979. [Google Scholar] [CrossRef] [PubMed]
  36. Grassberger, P.; Procaccia, L. Characterization of strange attractors. Phys. Rev. Lett. 1983, 50, 346–349. [Google Scholar] [CrossRef]
  37. Wolf, A.; Swift, J.B.; Swinney, H.L.; Vastano, J.A. Determining Lyapunov exponents from a time series. Physica D 1985, 16, 285–317. [Google Scholar] [CrossRef]
  38. Gao, J.; Cao, Y.; Hu, J. Multiscale Analysis of Complex Time Series; Wiley and Sons Interscience: Hoboken, NJ, USA, 2007; 368p. [Google Scholar]
  39. Chelani, A.; Devotta, S. Nonlinear analysis and prediction of coarse particulate matter concentration in ambient air. J. Air Waste Manag. Asoc. 2006, 56, 78–84. [Google Scholar] [CrossRef]
  40. Horna Mercedes, J.; Dionicio Vereau, J.; Martínez Zocón, R.; Zavaleta Quipuscoa, A.; Brenis Delgado, Y. Dinámica Simbólica y Algunas Aplicaciones. Sel. Matemáticas 2016, 3, 101–106. [Google Scholar] [CrossRef]
  41. Kantz, H.; Schreiber, T. Nonlinear Time Series Analysis, 2nd ed.; Cambridge University Press: Cambridge, UK, 2004; 387p. [Google Scholar]
  42. Sivakumar, B.; Wallender, W.; Horwath, W.; Mitchell, J. Nonlinear deterministic analysis of air pollution dynamics in a rural and agricultural setting. Adv. Complex Syst. 2007, 10, 581–597. [Google Scholar] [CrossRef]
  43. Lee, C.K.; Lin, S.C. Chaos in air pollutant concentration (APC) time series. Aerosol Air Qual. Res. 2008, 8, 381–391. [Google Scholar] [CrossRef]
  44. Kumar, U.; Prakash, A.; Jain, V.K. Characterization of chaos in air pollutants: A Volterra-Wiener-Korenberg series and numerical titration approach. Atmos. Environ. 2008, 42, 1537–1551. [Google Scholar] [CrossRef]
  45. Fraser, A.M.; Swinney, H.L. Independent coordinates for strange attractors from mutual information. Phys. Rev. A 1986, 33, 1134–1140. [Google Scholar] [CrossRef]
  46. Malyutina, E.I.; Shiryaev, V.I. Time Series Forecasting Using Nonlinear Dynamic Methods and Identification of Deterministic Chaos. Procedia Comput. Sci. 2014, 31, 1022–1103. [Google Scholar] [CrossRef]
  47. Salini, G.; Pérez, P. A study of the dynamic behavior of fine particulate matter in Santiago, Chile. Aerosol Air Qual. Res. 2015, 15, 154–165. [Google Scholar] [CrossRef]
  48. Pacheco, P.R.; Parodi, M.C.; Mera, E.M.; Salini, G.A. Variables meteorológicas y niveles de concentración de material particulado de 10 μm en Andacollo, Chile: Un estudio de dispersión y entropías. Inf. Tecnológica 2020, 31, 171–182. [Google Scholar] [CrossRef]
  49. Pacheco, P.R.; Salini, G.A.; Mera, E.M. Entropía y Neguentropía: Una aproximación al proceso de difusión de contaminantes y su sostenibilidad. Rev. Int. Contam. Ambient. 2021, 37, 167–185. [Google Scholar] [CrossRef]
  50. Salini, G.A.; Pacheco, P.R.; Mera, E.; Parodi, M.C. Probable Relationship between COVID-19, Pollutants and Meteorology: A Case Study at Santiago, Chile. Aerosol Air Qual. Res. 2021, 21, 200434. [Google Scholar] [CrossRef]
  51. Pacheco, P.; Mera, E.; Salini, G. Urban Densification Effect on Micrometeorology in Santiago, Chile: A Comparative Study Based on Chaos Theory. Sustainability 2022, 14, 2845. [Google Scholar] [CrossRef]
  52. Norazian, M.N.; Shruki, Y.A.; Azam, R.M.; Mustafa Al Bakri, A.M. Estimation of missing values in air pollution data using single imputation techniques. ScienceAsia 2008, 34, 341–345. [Google Scholar] [CrossRef]
  53. Junninen, H.; Niska, H.; Tuppurainen, K.; Ruuskanen, J.; Kolehmainen, M. Methods for imputation of missing values in air quality data sets. Atmos. Environ. 2004, 38, 2895–2907. [Google Scholar] [CrossRef]
  54. Asa, E.; Saafi, M.; Membah, J.; Billa, A. Comparison of linear and nonlinear Kriging methods for characterization and interpolation of soil data. J. Comput. Civil Eng. 2012, 26, 11–18. [Google Scholar] [CrossRef]
  55. Emery, X. Simple and ordinary multigaussian Kriging for estimating recoverable reserves. Math. Geol. 2005, 37, 295–319. [Google Scholar] [CrossRef]
  56. Kyriakidis, P.; Journel, A. Geostatistical space-time models: A review. Math. Geol. 1999, 6, 651–684. [Google Scholar] [CrossRef]
  57. Rutland, J.; Garreaud, R. Meteorological air pollution for Santiago, Chile: Towards an objective episode forecasting. Environ. Monit. Assess. 1995, 34, 223–244. [Google Scholar] [CrossRef] [PubMed]
  58. Rubio, M.A.; Fuenzalida, I.; Salinas, E.; Lissi, E.; Kurtenbach, R.; Wiesen, P. Carbon monoxide and carbon dioxide concentrations in Santiago de Chile associated with traffic emissions. Environ. Monit. Assess. 2010, 162, 209–217. [Google Scholar] [CrossRef] [PubMed]
  59. Livert Aquino, F.; Gainza, X. Understanding Density in an Uneven City, Santiago de Chile:Implications for Social and Environmental Sustainability. Sustainability 2014, 6, 5876–5897. [Google Scholar] [CrossRef]
  60. Vicuña, M. Density and its Effects on the Spatial Transformation of the Contemporary City: Five Typologies to Understand Residential Densification in Santiago Metropolitan Area. Revista 180 2020, 45, 112–126. [Google Scholar] [CrossRef]
  61. Saaroni, H.; Baruch Ziv, B. Estimating the Urban Heat Island Contribution to Urban and Rural Air Temperature Differences over Complex Terrain: Application to an Arid City. J. Appl. Meteorol. Climatol. 2010, 49, 2159–2166. [Google Scholar] [CrossRef]
  62. Tursilowati, L.; Sumantyo, J.T.S.; Kuze, H.; Adiningsih, E.S. The integrated wrf/urban modeling system and its application to monitoring urban heat island in Jakarta, indonesia. J. Urban Environ. Eng. 2012, 6, 1–9. [Google Scholar] [CrossRef]
  63. Lee, S.H.; Kim, S.W.; Angevine, W.; Bianco, L.; McKeen, S.; Senff, C.; Trainer, M.; Tucker, S.; Zamora, R. Evaluation of urban surface parameterizations in the WRF model using measurements during the Texas air quality study 2006 field campaign. Atmos. Chem. Phys. 2011, 11, 2127–2143. [Google Scholar] [CrossRef]
  64. Chen, Y.-C.; Fröhlich, D.; Matzarakis, A.; Ping Lin, T. Urban Roughness Estimation Based on Digital Building Models for Urban Wind and Thermal Condition Estimation—Application of the SkyHelios Model. Atmosphere 2017, 8, 247. [Google Scholar] [CrossRef]
  65. Vergara, J.E. Verticalización. La Edificación en Altura en la Región Metropolitana de Santiago (1990–2014), Revista INVI 2017, 32, 9–49. [Google Scholar]
  66. Craig, S.; Grinham, J. Breathing walls: The design of porous materials for heat exchange and decentralized ventilation. Energy Build. 2017, 149, 246–259. [Google Scholar] [CrossRef]
Figure 1. The described process is repeated for the six variables measured in each of the twelve locations, giving a total of 72 time series (each series contains 28,463 data).
Figure 1. The described process is repeated for the six variables measured in each of the twelve locations, giving a total of 72 time series (each series contains 28,463 data).
Atmosphere 14 00357 g001
Figure 2. Geographical distribution of the monitoring stations for this study (yellow stars network) and high construction population density (red color for the network of areas with the highest population density).
Figure 2. Geographical distribution of the monitoring stations for this study (yellow stars network) and high construction population density (red color for the network of areas with the highest population density).
Atmosphere 14 00357 g002
Figure 3. Distribution of entropies according to heights and periods of the meteorological variables.
Figure 3. Distribution of entropies according to heights and periods of the meteorological variables.
Atmosphere 14 00357 g003
Figure 4. Distribution of entropies according to heights and periods of the pollutants.
Figure 4. Distribution of entropies according to heights and periods of the pollutants.
Atmosphere 14 00357 g004
Figure 5. Height and CK according to the various communities analyzed. Series 1 period 2010–2013 and Series 2 period 2017–2020.
Figure 5. Height and CK according to the various communities analyzed. Series 1 period 2010–2013 and Series 2 period 2017–2020.
Atmosphere 14 00357 g005
Figure 6. Approximate modeling of the CK entropy ratio according to the height of each monitoring station and the two study periods. Series 3 and 4 are the experimental data from the periods 2010–2013 and 2017–2020, respectively.
Figure 6. Approximate modeling of the CK entropy ratio according to the height of each monitoring station and the two study periods. Series 3 and 4 are the experimental data from the periods 2010–2013 and 2017–2020, respectively.
Atmosphere 14 00357 g006
Figure 7. The approximate modeling of the CK entropy ratio according to the study periods, 2010–2013 (Series 1) and 2017–2020 (Series 2) with the domain of heights outside the range of measurements.
Figure 7. The approximate modeling of the CK entropy ratio according to the study periods, 2010–2013 (Series 1) and 2017–2020 (Series 2) with the domain of heights outside the range of measurements.
Atmosphere 14 00357 g007
Figure 8. The upper figure represents the period 2010–2013, which is considered the beginning of the intensive urban densification process, mainly due to high-rise buildings. The lower figure shows the result of the continuity and increase in densification (period 2017–2020). Δh = h − 469 (m).
Figure 8. The upper figure represents the period 2010–2013, which is considered the beginning of the intensive urban densification process, mainly due to high-rise buildings. The lower figure shows the result of the continuity and increase in densification (period 2017–2020). Δh = h − 469 (m).
Atmosphere 14 00357 g008
Table 1. Presents the entropies of the meteorological variables (MVs) for the two periods (2010–2013 and 2017–2020) according to the same heights.
Table 1. Presents the entropies of the meteorological variables (MVs) for the two periods (2010–2013 and 2017–2020) according to the same heights.
Stationm.a.s.l. (m)Series 1Series 2
2010–20132017–2020
SK,MV (1/h)SK,MV (1/h)
EML7841.3341.284
EMM7091.4521.205
EMO4691.1970.993
EMS6981.2891.250
EMV4851.2020.994
EMN5701.2621.145
Table 2. It presents the entropies of the pollutants (Ps) in the two periods of the study (2010–2013 and 2017–2020) carried out in the same locations according to their own heights.
Table 2. It presents the entropies of the pollutants (Ps) in the two periods of the study (2010–2013 and 2017–2020) carried out in the same locations according to their own heights.
Stationm.a.s.l. (m)Series 1Series 2
2010–20132017–2020
SK,P (1/h)SK,P (1/h)
EML7841.5421.577
EMM7091.5501.406
EMO4691.2101.630
EMS6981.3771.702
EMV4851.4311.220
EMN5701.5311.479
Table 3. Percentage variations of the CK.
Table 3. Percentage variations of the CK.
h (m.a.s.l.)2010–20132017–2020% Variation
CK1 (Series 1)CK2 (Series 2)
784 (EML)0.8650.8146
709 (EMM)0.9370.8577
485 (EMV)0.8340.8153
570 (EMN)0.8240.7747
698 (EMS)0.9360.73422
469 (EMO)0.9890.60938
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pacheco, P.; Mera, E.; Fuentes, V.; Parodi, C. Initial Conditions and Resilience in the Atmospheric Boundary Layer of an Urban Basin. Atmosphere 2023, 14, 357. https://doi.org/10.3390/atmos14020357

AMA Style

Pacheco P, Mera E, Fuentes V, Parodi C. Initial Conditions and Resilience in the Atmospheric Boundary Layer of an Urban Basin. Atmosphere. 2023; 14(2):357. https://doi.org/10.3390/atmos14020357

Chicago/Turabian Style

Pacheco, Patricio, Eduardo Mera, Voltaire Fuentes, and Carolina Parodi. 2023. "Initial Conditions and Resilience in the Atmospheric Boundary Layer of an Urban Basin" Atmosphere 14, no. 2: 357. https://doi.org/10.3390/atmos14020357

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop