Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (72)

Search Parameters:
Keywords = Pontryagin’s minimum principle

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 2973 KiB  
Article
Dynamic Analysis and Optimal Control of the Spread of Tungro Virus Disease in Rice Plants Considering Refugia Planting and Pesticide Application
by Rika Amelia, Nursanti Anggriani, Asep K. Supriatna and Noor Istifadah
Mathematics 2024, 12(24), 3979; https://doi.org/10.3390/math12243979 - 18 Dec 2024
Viewed by 265
Abstract
One of the main obstacles in rice cultivation is tungro disease, caused by Rice Tungro Spherical Virus (RTSV) and Rice Tungro Bacilliform Virus (RTBV), which are transmitted by green leafhopper vectors (Nephotettix virescens). This disease can be controlled by using pesticides [...] Read more.
One of the main obstacles in rice cultivation is tungro disease, caused by Rice Tungro Spherical Virus (RTSV) and Rice Tungro Bacilliform Virus (RTBV), which are transmitted by green leafhopper vectors (Nephotettix virescens). This disease can be controlled by using pesticides and refugia plants. Excessive use of pesticides can have negative impacts and high costs, so it is necessary to control the use of pesticides. In this study, a mathematical model of the spread of tungro virus disease in rice plants was developed by considering the characteristics of the virus, the presence of green leafhoppers and natural enemies, refugia planting, and pesticide use. From this model, dynamic and sensitivity analyses were carried out, and the optimal control theory was searched using the Pontryagin minimum principle. The analysis results showed three equilibriums: two non-endemic equilibriums (when plant and vector populations exist and when plant, vector, and natural enemy populations exist) and one endemic equilibrium. The non-endemic equilibrium will be asymptotically stable locally if R0<1. At the same time, the parameters that greatly influence the spread of this disease are parameters μ, μ2, and ϕ for local sensitivity analysis and α, a, β, b, ϕ, and μ2 for global sensitivity analysis. The results of the numerical simulation show that control using combined control is more effective in reducing the intensity of the spread of tungro disease in rice plants than control in the form of planting refugia plants as a source of food for natural enemies. The use of pesticides is sufficient for only four days, so the costs incurred are quite effective in controlling the spread of this disease. Full article
(This article belongs to the Special Issue Mathematical Methods and Models in Epidemiology)
Show Figures

Figure 1

Figure 1
<p>Flow diagram of the spread of tungro virus disease in rice plants, taking into account the factors of planting refugia plants and applying pesticides (development of the previous model [<a href="#B14-mathematics-12-03979" class="html-bibr">14</a>,<a href="#B15-mathematics-12-03979" class="html-bibr">15</a>,<a href="#B16-mathematics-12-03979" class="html-bibr">16</a>,<a href="#B17-mathematics-12-03979" class="html-bibr">17</a>]).</p>
Full article ">Figure 2
<p>Schematic diagram of mathematical modeling of the spread of tungro virus disease in rice plants by considering the factors of planting refugia plants and applying pesticides (development of the previous model [<a href="#B14-mathematics-12-03979" class="html-bibr">14</a>,<a href="#B15-mathematics-12-03979" class="html-bibr">15</a>,<a href="#B16-mathematics-12-03979" class="html-bibr">16</a>,<a href="#B17-mathematics-12-03979" class="html-bibr">17</a>]).</p>
Full article ">Figure 3
<p>Population dynamics of tungro disease when <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>&lt;</mo> <mn>1</mn> </mrow> </semantics></math>: (<b>a</b>) vector; (<b>b</b>) plant.</p>
Full article ">Figure 4
<p>Population dynamics of tungro disease when <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>&gt;</mo> <mn>1</mn> </mrow> </semantics></math>: (<b>a</b>) vector; (<b>b</b>) plant.</p>
Full article ">Figure 5
<p>Pesticide-controlled and uncontrolled populations: (<b>a</b>) rice plants infected with RTSV; (<b>b</b>) rice plants infected with RTSV+RTBV; (<b>c</b>) green planthopper vector infected with RTSV; (<b>d</b>) green planthopper vector infected with RTSV+RTBV.</p>
Full article ">Figure 6
<p>Optimal control of pesticide application.</p>
Full article ">Figure 7
<p>Natural enemy-controlled and uncontrolled populations: (<b>a</b>) rice plants infected with RTSV; (<b>b</b>) rice plants infected with RTSV+RTBV; (<b>c</b>) green planthopper vector infected with RTSV; (<b>d</b>) green planthopper vector infected with RTSV+RTBV.</p>
Full article ">Figure 8
<p>Optimal control of use of natural enemies.</p>
Full article ">Figure 9
<p>Population resulting from a combination of pesticide control with natural enemies and no control: (<b>a</b>) rice plants infected with RTSV; (<b>b</b>) rice plants infected with RTSV+RTBV; (<b>c</b>) green planthopper vector infected with RTSV; (<b>d</b>) green planthopper vector infected with RTSV+RTBV.</p>
Full article ">Figure 9 Cont.
<p>Population resulting from a combination of pesticide control with natural enemies and no control: (<b>a</b>) rice plants infected with RTSV; (<b>b</b>) rice plants infected with RTSV+RTBV; (<b>c</b>) green planthopper vector infected with RTSV; (<b>d</b>) green planthopper vector infected with RTSV+RTBV.</p>
Full article ">Figure 10
<p>Optimal control of combined use of pesticides and natural enemies.</p>
Full article ">Figure 11
<p>Population from comparison of each control: (<b>a</b>) rice plants infected with RTSV; (<b>b</b>) rice plants infected with RTSV+RTBV; (<b>c</b>) green planthopper vector infected with RTSV; (<b>d</b>) green planthopper vector infected with RTSV+RTBV.</p>
Full article ">
26 pages, 11186 KiB  
Article
Dynamic Response Control Strategy for Parallel Hybrid Ships Based on PMP-HMPC
by Enzhe Song, Zhijiang Liu, Chong Yao, Xiaojun Sun, Xuchang Yang and Minghui Bao
Processes 2024, 12(11), 2564; https://doi.org/10.3390/pr12112564 - 16 Nov 2024
Viewed by 455
Abstract
With increasingly stringent emission regulations, various clean fuel engines, electric propulsion systems, and renewable energy sources have been demonstratively applied in marine power systems. The development of control strategies that can effectively and efficiently coordinate the operation of multiple energy sources has become [...] Read more.
With increasingly stringent emission regulations, various clean fuel engines, electric propulsion systems, and renewable energy sources have been demonstratively applied in marine power systems. The development of control strategies that can effectively and efficiently coordinate the operation of multiple energy sources has become a key research focus. This study uses a modular modeling method to establish a system simulation model for a parallel hybrid ship with a natural gas engine (NGE) as the prime mover, and designs an energy management control strategy that can run in real time. The strategy is based on Pontryagin’s minimum principle (PMP) for power allocation, and is supplemented by a hybrid model predictive control (HMPC) method for speed-tracking control of the power system. Finally, the designed strategy is evaluated. Through simulation and hardware-in-the-loop (HIL) experimental validation, results compared with the Rule-based strategy indicate that under the given conditions, the SOC final value deviation from the initial value is reduced from 11.5% (in the reference strategy) to 0.39%. The system speed error integral is significantly lower at 39.06, compared to 2264.67 in the reference strategy. While gas consumption increased slightly by 2.4%, emissions were reduced by 3.2%. Full article
(This article belongs to the Section Energy Systems)
Show Figures

Figure 1

Figure 1
<p>Hybrid power testing apparatus.</p>
Full article ">Figure 2
<p>Hybrid power ship topology.</p>
Full article ">Figure 3
<p>Motor characteristics.</p>
Full article ">Figure 4
<p>Loading Conditions.</p>
Full article ">Figure 5
<p>Energy Management Strategy Framework.</p>
Full article ">Figure 6
<p>NGE Friction Torque and Thermal Efficiency Curves.</p>
Full article ">Figure 7
<p>(<b>a</b>) Linearization of the NGE External Characteristic; (<b>b</b>) Linearization of the PMSM External Characteristic.</p>
Full article ">Figure 8
<p>(<b>a</b>) Gas Consumption Rate Linearization; (<b>b</b>) NGE Emissions Linearization.</p>
Full article ">Figure 9
<p>Results of Battery Power Linearization.</p>
Full article ">Figure 10
<p>(<b>a</b>) Speed Tracking of the Power System; (<b>b</b>) Torque Distribution of the NGE; (<b>c</b>) Torque Distribution of the PMSM; (<b>d</b>) SOC Variation.</p>
Full article ">Figure 10 Cont.
<p>(<b>a</b>) Speed Tracking of the Power System; (<b>b</b>) Torque Distribution of the NGE; (<b>c</b>) Torque Distribution of the PMSM; (<b>d</b>) SOC Variation.</p>
Full article ">Figure 11
<p>Results of Speed Error Integration.</p>
Full article ">Figure 12
<p>Variation of Gas Consumption Under Three Strategies.</p>
Full article ">Figure 13
<p>Variation of Emissions Under Three Strategies.</p>
Full article ">Figure 14
<p>Radar Chart of Three Control Strategies.</p>
Full article ">Figure 15
<p>HIL test system.</p>
Full article ">Figure 16
<p>(<b>a</b>) Torque distribution; (<b>b</b>) Speed Variation; (<b>c</b>) SOC.</p>
Full article ">
37 pages, 4566 KiB  
Article
Aperiodic Optimal Chronotherapy in Simple Compartment Tumour Growth Models Under Circadian Drug Toxicity Conditions
by Byron D. E. Tzamarias, Annabelle Ballesta and Nigel John Burroughs
Mathematics 2024, 12(22), 3516; https://doi.org/10.3390/math12223516 - 11 Nov 2024
Viewed by 479
Abstract
Cancer cells typically divide with weaker synchronisation with the circadian clock than normal cells, with the degree of decoupling increasing with tumour maturity. Chronotherapy exploits this loss of synchronisation, using drugs with circadian-clock-dependent activity and timed infusion to balance the competing demands of [...] Read more.
Cancer cells typically divide with weaker synchronisation with the circadian clock than normal cells, with the degree of decoupling increasing with tumour maturity. Chronotherapy exploits this loss of synchronisation, using drugs with circadian-clock-dependent activity and timed infusion to balance the competing demands of reducing toxicity toward normal cells that display physiological circadian rhythms and of efficacy against the tumour. We analysed optimal chronotherapy for one-compartment nonlinear tumour growth models that were no longer synchronised with the circadian clock, minimising a cost function with a periodically driven running cost accounting for the circadian drug tolerability of normal cells. Using Pontryagin’s Minimum Principle (PMP), we show, for drugs that either increase the cell death rate or kill dividing cells, that optimal solutions are aperiodic bang–bang solutions with two switches per day, with the duration of the daily drug administration increasing as treatment progresses; for large tumours, optimal therapy can in fact switch mid treatment from aperiodic to continuous treatment. We illustrate this with tumours grown under logistic and Gompertz dynamics conditions; for logistic growth, we categorise the different types of solutions. Singular solutions can be applicable for some nonlinear tumour growth models if the per capita growth rate is convex. Direct comparison of the optimal aperiodic solution with the optimal periodic solution shows the former presents reduced toxicity whilst retaining similar efficacy against the tumour. We only found periodic solutions with a daily period in one-compartment exponential growth models, whilst models incorporating nonlinear growth had generic aperiodic solutions, and linear multi-compartments appeared to have long-period (weeks) periodic solutions. Our results suggest that chronotherapy-based optimal solutions under a harmonic running cost are not typically periodic infusion schedules with a 24 h period. Full article
(This article belongs to the Section Mathematical Biology)
Show Figures

Figure 1

Figure 1
<p>Graphical representation of trajectory dynamics, which are admissible under PMP, in the <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>−</mo> <mi>w</mi> </mrow> </semantics></math> plane. Trajectories start on the initialisation line (tumour size <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <msub> <mi>N</mi> <mn>0</mn> </msub> </mrow> </semantics></math>), i.e., the black dashed vertical line, and end on the termination line <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>=</mo> <mi>r</mi> <mi>N</mi> </mrow> </semantics></math>, (red line). The <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>−</mo> <mi>w</mi> </mrow> </semantics></math> plane is divided into three zones: zone 1 (<math display="inline"><semantics> <mrow> <mi>w</mi> <mo>&lt;</mo> <mo movablelimits="true" form="prefix">min</mo> <mo>(</mo> <mi>h</mi> <mo>)</mo> </mrow> </semantics></math>), with <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>; zone 2 (<math display="inline"><semantics> <mrow> <mo movablelimits="true" form="prefix">min</mo> <mo>(</mo> <mi>h</mi> <mo>)</mo> <mo>≤</mo> <mi>w</mi> <mo>&lt;</mo> <mo movablelimits="true" form="prefix">max</mo> <mo>(</mo> <mi>h</mi> <mo>)</mo> </mrow> </semantics></math>), the switching zone, where <span class="html-italic">u</span> is either 0 or 1; and zone 3 (<math display="inline"><semantics> <mrow> <mi>w</mi> <mo>≥</mo> <mo movablelimits="true" form="prefix">max</mo> <mo>(</mo> <mi>h</mi> <mo>)</mo> </mrow> </semantics></math>), with <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>. Finally, the nullclines are shown (region dependent), with <span class="html-italic">w</span>-nullcline <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> in blue, <span class="html-italic">N</span>-nullclines <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> in orange, and <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <msub> <mi>N</mi> <mo>*</mo> </msub> </mrow> </semantics></math> in magenta. In zone 1, the orange vertical line <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> is both a <span class="html-italic">w</span> and <span class="html-italic">N</span> nullcline. A trajectory is illustrated in green starting from <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <msub> <mi>N</mi> <mn>0</mn> </msub> </mrow> </semantics></math> (the dashed black line) and terminating when it reaches the (red) termination line <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>=</mo> <mi>r</mi> <mi>N</mi> </mrow> </semantics></math>. Trajectories can only occur in the upper quadrant. Sketch is based on the Gompertz model with a carrying capacity of <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mo>*</mo> </msub> <mo>=</mo> <mn>8</mn> <mo>·</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math> cells.</p>
Full article ">Figure 2
<p>Trajectory solutions in the <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>−</mo> <mi>w</mi> </mrow> </semantics></math> plane for the logistic model. There are 3 cases depending on the initial tumours: (<b>a</b>) large initial tumour size, <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>&gt;</mo> <msub> <mi>N</mi> <mi>b</mi> </msub> </mrow> </semantics></math>; (<b>b</b>) intermediate initial tumours, <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>b</mi> </msub> <mo>&gt;</mo> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>&gt;</mo> <msub> <mi>N</mi> <mi>a</mi> </msub> </mrow> </semantics></math>; (<b>c</b>) small initial tumours <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>&lt;</mo> <msub> <mi>N</mi> <mi>a</mi> </msub> </mrow> </semantics></math>. Trajectories are colour coded by type, classified in <a href="#mathematics-12-03516-f003" class="html-fig">Figure 3</a>. The switching zone boundaries, <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mn>1</mn> <mo>−</mo> <mi>α</mi> </mrow> <mi>d</mi> </mfrac> </mstyle> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mn>1</mn> <mo>+</mo> <mi>α</mi> </mrow> <mi>d</mi> </mfrac> </mstyle> </mrow> </semantics></math>, are indicated by black horizontal lines. The vertical solid lines (<math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <msub> <mi>N</mi> <mi>a</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <msub> <mi>N</mi> <mi>b</mi> </msub> </mrow> </semantics></math>) indicate the intersection of the termination line with the lower and upper switching bounds, respectively. Each trajectory starts on the intialisation line (<math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <msub> <mi>N</mi> <mn>0</mn> </msub> </mrow> </semantics></math>), the black dashed one, and terminates on the termination line (<math display="inline"><semantics> <mrow> <mi>w</mi> <mo>=</mo> <mi>r</mi> <mi>N</mi> </mrow> </semantics></math>), shown in thick black lines. Grey arrows represent the flow of solutions in zones 1 and 3 respectively. Parameter values are <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>0.0625</mn> </mrow> </semantics></math> per day, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>0.125</mn> </mrow> </semantics></math> per day, <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>50</mn> <mo>·</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>8</mn> </mrow> </msup> </mrow> </semantics></math> days per cell, <math display="inline"><semantics> <mrow> <mi>K</mi> <mo>=</mo> <msup> <mn>10</mn> <mn>8</mn> </msup> </mrow> </semantics></math> cells, and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Classification of solutions in the <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>−</mo> <mi>w</mi> </mrow> </semantics></math> plane for the logistic model for large tumours (case A, <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>&gt;</mo> <msub> <mi>N</mi> <mi>b</mi> </msub> </mrow> </semantics></math>). The switching zone boundaries are indicated by horizontal black lines. Possible trajectories start on the initialisation line (dashed, <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <msub> <mi>N</mi> <mn>0</mn> </msub> </mrow> </semantics></math>) and terminate on the termination line (thick black line lying close to the <span class="html-italic">w</span>-axis). Trajectories are classified as follows: <b>Type-0</b> (brown) starts in zone 1, traverses across zone 2, and ends in zone 3; the tumor size increases until the switching zone is reached, where the trajectories drift towards the termination line, reaching termination only under continuous MTD in zone 3. <b>Type-1</b> (red) trajectories are similar to Type-0 but start in zone 1 and traverse to and end in zone 2; the tumor size increases until the switching zone is reached, where the trajectories drift towards the termination line. <b>Type-2</b> (blue) trajectories start and end in zone 2, with <span class="html-italic">N</span> increasing beyond <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math> (there exists <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mi>T</mi> <mo>]</mo> </mrow> </semantics></math> s.t. <math display="inline"><semantics> <mrow> <mi>N</mi> <mrow> <mo>(</mo> <mi>s</mi> <mo>)</mo> </mrow> <mo>&gt;</mo> <msub> <mi>N</mi> <mn>0</mn> </msub> </mrow> </semantics></math>); despite starting in the switching zone with daily drug administration, these trajectories take a “detour“ away from the termination line to larger tumour sizes. <b>Type-3</b> trajectories (black) start and end in zone-2, with the tumor size always less than <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>; these trajectories are the most cost-efficient, with daily switching reducing drug toxicity, and a low final tumor size can be attained. <b>Type-4</b> (purple) trajectories start in zone-2 and end in zone-3; they are similar to type-3 trajectories except they shift to continuous MTD at the end of treatment;. <b>Type-5</b> (orange) trajectories start and end in zone-3; MTD is applied continuously, and the terminal tumor size approaches <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math> as <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>→</mo> <mn>0</mn> </mrow> </semantics></math>. Parameter values: <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.07</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>K</mi> <mo>=</mo> <msup> <mn>10</mn> <mn>8</mn> </msup> </mrow> </semantics></math> cells, <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>5000</mn> <mo>·</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>8</mn> </mrow> </msup> </mrow> </semantics></math> days per cell, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>0.125</mn> </mrow> </semantics></math> per day, and <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>0.025</mn> </mrow> </semantics></math> per day. Parameter values were chosen to illustrate the types of optimal solutions.</p>
Full article ">Figure 4
<p>Daily drug infusion and tumour decay for a type 3 solution to the logistic model. (<b>a</b>) Classification of solutions (colour coding by type as in <a href="#mathematics-12-03516-f003" class="html-fig">Figure 3</a>); there is no type 0 for these parameter values and the initial tumour size. (<b>b</b>) Daily drug administration duration (hours). (<b>c</b>) Daily relative reduction in tumor size (relative to the beginning of the day) throughout the treatment. The parameter values are the same as in <a href="#mathematics-12-03516-f003" class="html-fig">Figure 3</a> except for <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math>, with an initial tumor size of <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>=</mo> <msup> <mn>10</mn> <mn>7</mn> </msup> </mrow> </semantics></math> cells. Time horizon is <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>45</mn> </mrow> </semantics></math> days in panels (<b>b</b>,<b>c</b>).</p>
Full article ">Figure 5
<p>Logistic model solution costs and time horizons. (<b>a</b>) The dependence of the time horizon, (<b>b</b>) total cost, (<b>c</b>) terminal cost, and (<b>d</b>) running cost on the initial value of <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math>. Solution types are colour-coded as in <a href="#mathematics-12-03516-f003" class="html-fig">Figure 3</a>. The inset plots in panels (<b>a</b>,<b>b</b>) show the time horizon and cost as a function of <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> for Type-5 solutions. The initial tumor size and parameter values are given as in <a href="#mathematics-12-03516-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 6
<p>Details of cost and time horizon dependence on <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> for the logistic model. (<b>a</b>) The cost and (<b>b</b>) the time horizon for a narrow region of <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> values around the global minimum cost solution. The parameter values and the initial tumour size are the same as those in <a href="#mathematics-12-03516-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 7
<p>Details of <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> as termination is approached for 3 solutions with close <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> values (logistic model). The 3 cases correspond to (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>9.2410</mn> </mrow> </semantics></math> that terminates just before the next switching time, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>9.2364</mn> </mrow> </semantics></math> that terminates at the switching time, and (<b>c</b>) that reaches the termination line and terminates just after a switching event. Switching and termination points are represented with red and black marks, respectively, thick black line is the trajectory <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> and thin black curves are the function <math display="inline"><semantics> <mrow> <mi>h</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math>. The areas of the <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>−</mo> <mi>t</mi> </mrow> </semantics></math> plane where <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> <mo>&lt;</mo> <mi>h</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> <mo>&gt;</mo> <mi>h</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> are coloured white and pink, respectively. All trajectories reach the termination line when <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>≥</mo> <mi>h</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> </mrow> </semantics></math>, resulting in a discontinuity in the time horizon as <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> decreases past <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>9.2364</mn> </mrow> </semantics></math> (case (<b>b</b>)). Parameter values and the initial tumour size are the same as in <a href="#mathematics-12-03516-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 7 Cont.
<p>Details of <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> as termination is approached for 3 solutions with close <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> values (logistic model). The 3 cases correspond to (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>9.2410</mn> </mrow> </semantics></math> that terminates just before the next switching time, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>9.2364</mn> </mrow> </semantics></math> that terminates at the switching time, and (<b>c</b>) that reaches the termination line and terminates just after a switching event. Switching and termination points are represented with red and black marks, respectively, thick black line is the trajectory <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> and thin black curves are the function <math display="inline"><semantics> <mrow> <mi>h</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math>. The areas of the <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>−</mo> <mi>t</mi> </mrow> </semantics></math> plane where <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> <mo>&lt;</mo> <mi>h</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> <mo>&gt;</mo> <mi>h</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> are coloured white and pink, respectively. All trajectories reach the termination line when <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>≥</mo> <mi>h</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> </mrow> </semantics></math>, resulting in a discontinuity in the time horizon as <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> decreases past <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>9.2364</mn> </mrow> </semantics></math> (case (<b>b</b>)). Parameter values and the initial tumour size are the same as in <a href="#mathematics-12-03516-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 8
<p>Construction of solutions with multiple termination line crossings (logistic model). (<b>a</b>) A segment of a trajectory in the <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>−</mo> <mi>w</mi> </mrow> </semantics></math> plane that intersects the termination line three times. Trajectory components are colour coded, before the first crossing of the termination line (black), after first crossing and before second (orange), after second crossing (green). Termination line is shown (black) and crossings indicated by dots coloured as the prior trajectory segment. Switching points are shown (red). Trajectory starts at <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.1</mn> <mo>·</mo> <msup> <mn>10</mn> <mn>8</mn> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>5.699</mn> </mrow> </semantics></math> and can terminate on the first, second or third crossing giving 3 horizon times for that <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math>. (<b>b</b>) The time horizon as a function of <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> over range <math display="inline"><semantics> <mrow> <mn>7.2</mn> <mo>≤</mo> <msub> <mi>w</mi> <mn>0</mn> </msub> <mo>≤</mo> <mn>7.3</mn> </mrow> </semantics></math>. Points that correspond to trajectories that cross the termination line once, twice and three times are coloured in black, orange and green respectively. The parameter values and the initial tumour size are the same as those in <a href="#mathematics-12-03516-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 9
<p>The cost for trajectories that cross the termination line up to three times in the logistic model. (<b>a</b>) The dependence of the cost on <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math>, in a region where the cost takes its minimum value. (<b>b</b>) The cost of the same solutions as in (<b>a</b>) shown as a function of the time horizon, <span class="html-italic">T</span>. Trajectories are coloured by number of crossings—1 black, 2 orange, and 3 green. The parameter values and the initial tumour size are the same as in <a href="#mathematics-12-03516-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 10
<p>Gompertz model solution costs and time horizons. (<b>a</b>) The dependence of the time horizon, (<b>b</b>) total cost, (<b>c</b>) terminal cost, and (<b>d</b>) running cost on the initial value of <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math>. Colours correspond to the solution type classification in <a href="#mathematics-12-03516-f003" class="html-fig">Figure 3</a>. The inset plots in panels (<b>a</b>,<b>b</b>) show the time horizon and cost as a function of <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> for type-5 solutions. Parameter values: <math display="inline"><semantics> <mrow> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>ε</mi> <mi>λ</mi> </mfrac> </mstyle> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mn>1</mn> <mn>3</mn> </mfrac> </mstyle> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.025</mn> </mrow> </semantics></math> per day, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>0.1373</mn> </mrow> </semantics></math>, per day <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>5000</mn> <mo>·</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>8</mn> </mrow> </msup> </mrow> </semantics></math> days per cell, and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math>. The initial tumor size is <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.1</mn> <mo>·</mo> <msup> <mn>10</mn> <mn>8</mn> </msup> </mrow> </semantics></math> cells. Parameter values were selected to illustrate behaviour of the optimal solutions.</p>
Full article ">Figure 10 Cont.
<p>Gompertz model solution costs and time horizons. (<b>a</b>) The dependence of the time horizon, (<b>b</b>) total cost, (<b>c</b>) terminal cost, and (<b>d</b>) running cost on the initial value of <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math>. Colours correspond to the solution type classification in <a href="#mathematics-12-03516-f003" class="html-fig">Figure 3</a>. The inset plots in panels (<b>a</b>,<b>b</b>) show the time horizon and cost as a function of <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> for type-5 solutions. Parameter values: <math display="inline"><semantics> <mrow> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mi>ε</mi> <mi>λ</mi> </mfrac> </mstyle> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mn>1</mn> <mn>3</mn> </mfrac> </mstyle> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.025</mn> </mrow> </semantics></math> per day, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>0.1373</mn> </mrow> </semantics></math>, per day <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>5000</mn> <mo>·</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>8</mn> </mrow> </msup> </mrow> </semantics></math> days per cell, and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math>. The initial tumor size is <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.1</mn> <mo>·</mo> <msup> <mn>10</mn> <mn>8</mn> </msup> </mrow> </semantics></math> cells. Parameter values were selected to illustrate behaviour of the optimal solutions.</p>
Full article ">Figure 11
<p>Cost of multiple termination-line-crossing solutions for the Gompertz model. (<b>a</b>) The dependence of the cost on <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> shown over the range <math display="inline"><semantics> <mrow> <mn>6.49</mn> <mo>≤</mo> <msub> <mi>w</mi> <mn>0</mn> </msub> <mo>≤</mo> <mn>6.52</mn> </mrow> </semantics></math>. Termination points that correspond to trajectories that satisfy the transversality condition at the first, second, or third crossing of the termination line are coloured in black, orange, and green, respectively. (<b>b</b>) The terminal value <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> </mrow> </semantics></math> against the time horizon for trajectories in panel (<b>a</b>). The control value upon termination, <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> </mrow> </semantics></math>, is shown by background shading; pink area is <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>; and white area is <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>. Termination points coloured as in panel (<b>a</b>). The parameters and the initial tumour size are the same as in <a href="#mathematics-12-03516-f010" class="html-fig">Figure 10</a>.</p>
Full article ">Figure 12
<p>Switching function and drug scheduling for a solution satisfying PMP. (<b>a</b>) The switching function <math display="inline"><semantics> <mrow> <mo>Φ</mo> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math>; green and red colors illustrate time periods where the control <span class="html-italic">u</span> is 1 and 0, respectively. The inset plot shows <math display="inline"><semantics> <mrow> <mo>Φ</mo> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> for the first 41 days of therapy. (<b>b</b>) the percentage of each day in which a drug is administered; orange bars relate to days in which the drug is chronomodulated, blue bars relate to days wherein MTD is administered continuously, and the horizontal line shows the respective percentage of the optimised periodic drug regimen. The inset plot illustrates the percentage for the first 41 days of therapy. The time horizon is 1 year, the number of dividing (<math display="inline"><semantics> <msub> <mi>N</mi> <mn>1</mn> </msub> </semantics></math>) and non-dividing (<math display="inline"><semantics> <msub> <mi>N</mi> <mn>2</mn> </msub> </semantics></math>) cancer cells at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> are <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mn>1</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>6.4376</mn> <mo>·</mo> <msup> <mn>10</mn> <mn>7</mn> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <mn>3.5624</mn> <mo>·</mo> <msup> <mn>10</mn> <mn>7</mn> </msup> </mrow> </semantics></math>; the growth rates of dividing and non-dividing cancer cells are <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0.1970</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.3560</mn> </mrow> </semantics></math>, respectively. The parameters that weight the terminal tumour size in the cost function are <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>52.5339</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>31.2068</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 13
<p>Tumour size for the PMPDR solution: (<b>a</b>) Number of dividing cancer cells (<math display="inline"><semantics> <msub> <mi>N</mi> <mn>1</mn> </msub> </semantics></math>) as a function of time, (<b>b</b>) number of non-dividing cancer cells (<math display="inline"><semantics> <msub> <mi>N</mi> <mn>2</mn> </msub> </semantics></math>) as a function of time. The inset plots in (<b>a</b>,<b>b</b>) show details of the time evolution of <math display="inline"><semantics> <msub> <mi>N</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>N</mi> <mn>2</mn> </msub> </semantics></math>, respectively.</p>
Full article ">Figure 14
<p>Example trajectories for a tumour with low (<math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.19</mn> </mrow> </semantics></math> per day, panels (<b>a</b>,<b>b</b>)) and high (<math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>2.41</mn> </mrow> </semantics></math> per day, panels (<b>c</b>,<b>d</b>)) susceptibility to the drug. Tumour evolution was assumed to be Gompertzian. (<b>a</b>,<b>c</b>) <span class="html-italic">N</span> as a function of time, and (<b>b</b>,<b>d</b>) <span class="html-italic">w</span> (dashed line) as a function of time. The periodic switching threshold <math display="inline"><semantics> <mrow> <mi>h</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> is shown in blue and the upper bound of the switching zone in black. <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>w</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> at switching and termination points are shown as black and red dots respectively. The inset panels show the time evolution of <span class="html-italic">N</span> and <span class="html-italic">w</span> over a longer time period. Parameters are <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.0792</mn> </mrow> </semantics></math> per day, <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>1500</mn> </mrow> </semantics></math> days/cell, <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>ε</mi> <mo>=</mo> <mn>77,293</mn> </mrow> </semantics></math> cells, and <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>1.865</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>10</mn> </msup> </mrow> </semantics></math> cells, and the drug-induced death rates are <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.19</mn> <mo>,</mo> </mrow> </semantics></math> and 2.41 per day for panels (<b>a</b>,<b>b</b>) and (<b>c</b>,<b>d</b>), respectively. Initial tumour size is <math display="inline"><semantics> <msup> <mn>10</mn> <mn>8</mn> </msup> </semantics></math> cells.</p>
Full article ">Figure 15
<p>Comparison of (unconstrained) optimal solutions (in light red) and optimal solutions constrained to be periodic (in black) for tumours with low susceptibility to the drug (Gompertz model). (<b>a</b>) Terminal tumour size versus <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>, (<b>b</b>) toxicity to normal cells (running cost) versus <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>, (<b>c</b>) treatment duration versus <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>, (<b>d</b>) total drug quantity administered versus <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>, (<b>e</b>) the fractional administration time versus <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>, and (<b>f</b>) percentage reduction of the running cost of unconstrained compared to constrained optimal solutions. Model parameters are the same as in <a href="#mathematics-12-03516-f014" class="html-fig">Figure 14</a>, with <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.19</mn> </mrow> </semantics></math> per day.</p>
Full article ">Figure 16
<p>Comparison of (unconstrained) optimal solutions (in light red), optimal constrained (periodic) solutions (asterisks), and continuous MTD (square) for tumours with high drug susceptibility (Gompertz model). (<b>a</b>) Terminal tumour size versus <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>, (<b>b</b>) toxicity to normal cells (running cost) versus <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>, (<b>c</b>) treatment duration versus <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>, (<b>d</b>) total drug administered versus <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>, (<b>e</b>) the fractional administration time versus <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>, and (<b>f</b>) percentage reduction of the running cost of specified treatments versus <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>. Model parameters as <a href="#mathematics-12-03516-f014" class="html-fig">Figure 14</a>, <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>2.41</mn> </mrow> </semantics></math> per day.</p>
Full article ">Figure A1
<p>Estimated computational errors of the switching and termination times for the logistic and the Gompertz models. We define as the absolute error of a trajectory the maximum of the quantities <math display="inline"><semantics> <mrow> <mo>∣</mo> <mi>h</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> <mo>−</mo> <mi>w</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> <mo>∣</mo> <mo>,</mo> <mo>.</mo> <mo>.</mo> <mo>.</mo> <mo>,</mo> <mo>∣</mo> <mi>h</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> <mo>)</mo> </mrow> <mo>−</mo> <mi>w</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> <mo>)</mo> </mrow> <mo>∣</mo> </mrow> </semantics></math> (where <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>,</mo> <mo>.</mo> <mo>.</mo> <mo>.</mo> <mo>,</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> </mrow> </semantics></math> are all switching times) and <math display="inline"><semantics> <mrow> <mo>∣</mo> <mi>r</mi> <mi>N</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>−</mo> <mi>w</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>∣</mo> </mrow> </semantics></math>, where <span class="html-italic">T</span> is the termination time. Similarly, the relative error of a trajectory is defined as the maximum of the quantities <math display="inline"><semantics> <mrow> <mo stretchy="false">∣</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mi>h</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> <mo>−</mo> <mi>w</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> </mrow> <mrow> <mo movablelimits="true" form="prefix">max</mo> <mo>(</mo> <mi>h</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> <mo>,</mo> <mi>w</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> <mo>)</mo> </mrow> </mfrac> </mstyle> <mo stretchy="false">∣</mo> <mo>,</mo> <mo>…</mo> <mo>,</mo> <mo stretchy="false">∣</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mi>h</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> <mo>)</mo> </mrow> <mo>−</mo> <mi>w</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> <mo>)</mo> </mrow> </mrow> <mrow> <mo movablelimits="true" form="prefix">max</mo> <mo>(</mo> <mi>h</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> <mo>)</mo> </mrow> <mo>,</mo> <mi>w</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> <mo>)</mo> </mrow> <mo>)</mo> </mrow> </mfrac> </mstyle> <mo stretchy="false">∣</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo stretchy="false">∣</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mi>r</mi> <mi>N</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>−</mo> <mi>w</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> </mrow> <mrow> <mo movablelimits="true" form="prefix">max</mo> <mo>(</mo> <mi>r</mi> <mi>N</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>,</mo> <mi>w</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>)</mo> </mrow> </mfrac> </mstyle> <mo stretchy="false">∣</mo> </mrow> </semantics></math>. For both models, the absolute and relative errors of 2394 trajectories were evaluated. Panels (<b>a</b>,<b>b</b>): absolute and relative errors in the logistic model; panels (<b>c</b>,<b>d</b>): absolute and relative errors in the Gompertz model.</p>
Full article ">Figure A1 Cont.
<p>Estimated computational errors of the switching and termination times for the logistic and the Gompertz models. We define as the absolute error of a trajectory the maximum of the quantities <math display="inline"><semantics> <mrow> <mo>∣</mo> <mi>h</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> <mo>−</mo> <mi>w</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> <mo>∣</mo> <mo>,</mo> <mo>.</mo> <mo>.</mo> <mo>.</mo> <mo>,</mo> <mo>∣</mo> <mi>h</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> <mo>)</mo> </mrow> <mo>−</mo> <mi>w</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> <mo>)</mo> </mrow> <mo>∣</mo> </mrow> </semantics></math> (where <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>,</mo> <mo>.</mo> <mo>.</mo> <mo>.</mo> <mo>,</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> </mrow> </semantics></math> are all switching times) and <math display="inline"><semantics> <mrow> <mo>∣</mo> <mi>r</mi> <mi>N</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>−</mo> <mi>w</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>∣</mo> </mrow> </semantics></math>, where <span class="html-italic">T</span> is the termination time. Similarly, the relative error of a trajectory is defined as the maximum of the quantities <math display="inline"><semantics> <mrow> <mo stretchy="false">∣</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mi>h</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> <mo>−</mo> <mi>w</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> </mrow> <mrow> <mo movablelimits="true" form="prefix">max</mo> <mo>(</mo> <mi>h</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> <mo>,</mo> <mi>w</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mn>1</mn> </msub> <mo>)</mo> </mrow> <mo>)</mo> </mrow> </mfrac> </mstyle> <mo stretchy="false">∣</mo> <mo>,</mo> <mo>…</mo> <mo>,</mo> <mo stretchy="false">∣</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mi>h</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> <mo>)</mo> </mrow> <mo>−</mo> <mi>w</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> <mo>)</mo> </mrow> </mrow> <mrow> <mo movablelimits="true" form="prefix">max</mo> <mo>(</mo> <mi>h</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> <mo>)</mo> </mrow> <mo>,</mo> <mi>w</mi> <mrow> <mo>(</mo> <msub> <mi>t</mi> <mi>q</mi> </msub> <mo>)</mo> </mrow> <mo>)</mo> </mrow> </mfrac> </mstyle> <mo stretchy="false">∣</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo stretchy="false">∣</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mi>r</mi> <mi>N</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>−</mo> <mi>w</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> </mrow> <mrow> <mo movablelimits="true" form="prefix">max</mo> <mo>(</mo> <mi>r</mi> <mi>N</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>,</mo> <mi>w</mi> <mo>(</mo> <mi>T</mi> <mo>)</mo> <mo>)</mo> </mrow> </mfrac> </mstyle> <mo stretchy="false">∣</mo> </mrow> </semantics></math>. For both models, the absolute and relative errors of 2394 trajectories were evaluated. Panels (<b>a</b>,<b>b</b>): absolute and relative errors in the logistic model; panels (<b>c</b>,<b>d</b>): absolute and relative errors in the Gompertz model.</p>
Full article ">Figure A2
<p>Daily drug dosage administered as a function of the death rate of cancer cells <span class="html-italic">d</span>. Each bar relates to a set of optimal solutions that minimize the cost (in terms of both time horizon and drug administration <span class="html-italic">u</span>) for a range of <span class="html-italic">r</span> and <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math> values. (<b>a</b>,<b>b</b>) illustrate results that are generated from the exponential and Gompertz tumour growth models, respectively. In the Gompertz case study, optimal solutions are non periodic; therefore, for each investigated solution, the average amount of drug that was administered daily was evaluated. The parameter values employed for the exponential model are <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.2</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>1.38</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>1.44</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>1.64</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>1.75</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>1.97</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.1</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.37</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.7</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.96</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.98</mn> </mrow> </semantics></math> cells/day; <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>0.1</mn> <mo>,</mo> <mn>10</mn> <mo>,</mo> <mn>100</mn> <mo>,</mo> <mn>1000</mn> <mo>,</mo> <mn>1500</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>0.9816</mn> </mrow> </semantics></math> cells/day; <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>; and <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>=</mo> <msup> <mn>10</mn> <mn>3</mn> </msup> <mo>,</mo> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mn>5</mn> </msup> <mo>,</mo> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mn>7</mn> </msup> <mo>,</mo> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mn>9</mn> </msup> </mrow> </semantics></math> cells. The parameter values employed for the Gompertz model are <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.39</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>1.58</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>1.63</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>1.7</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>1.8</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>1.9</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.1</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.2</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.3</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.4</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.5</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.6</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.7</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.8</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>2.9</mn> </mrow> </semantics></math> cells/day; <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>0.1</mn> <mo>,</mo> <mn>10</mn> <mo>,</mo> <mn>100</mn> <mo>,</mo> <mn>1000</mn> <mo>,</mo> <mn>1500</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>g</mi> <mo>=</mo> <mn>0.0792</mn> </mrow> </semantics></math> cells/day; <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>1.865</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>9</mn> </msup> </mrow> </semantics></math> cells; <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>77,293</mn> </mrow> </semantics></math> cells; <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>; and <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mn>0</mn> </msub> <mo>=</mo> <msup> <mn>10</mn> <mn>3</mn> </msup> <mo>,</mo> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mn>5</mn> </msup> <mo>,</mo> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mn>7</mn> </msup> <mo>,</mo> <mspace width="0.166667em"/> <msup> <mn>10</mn> <mn>9</mn> </msup> </mrow> </semantics></math> cells.</p>
Full article ">Figure A3
<p>The cost of optimal solutions as a function of <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> for the Gompertz model. Each panel corresponds to given values of <span class="html-italic">r</span> and <span class="html-italic">d</span>, each curve relates to a given value of <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>, and red points represent global minima. The vertical line indicates the value of <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> at the upper bound of the switching zone. (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>1500</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.19</mn> </mrow> </semantics></math> cells/day, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>1500</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>2.41</mn> </mrow> </semantics></math> cells/day, panel (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.19</mn> </mrow> </semantics></math> cells/day, and (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>2.41</mn> </mrow> </semantics></math> cells/day. All illustrated trajectories relate to the following parameter values: <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>1.865</mn> <mo>·</mo> <msup> <mn>10</mn> <mn>10</mn> </msup> </mrow> </semantics></math> cells, <math display="inline"><semantics> <mrow> <mi>ε</mi> <mo>=</mo> <mn>77,293</mn> </mrow> </semantics></math> cells, and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>.</p>
Full article ">Figure A3 Cont.
<p>The cost of optimal solutions as a function of <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> for the Gompertz model. Each panel corresponds to given values of <span class="html-italic">r</span> and <span class="html-italic">d</span>, each curve relates to a given value of <math display="inline"><semantics> <msub> <mi>N</mi> <mn>0</mn> </msub> </semantics></math>, and red points represent global minima. The vertical line indicates the value of <math display="inline"><semantics> <msub> <mi>w</mi> <mn>0</mn> </msub> </semantics></math> at the upper bound of the switching zone. (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>1500</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.19</mn> </mrow> </semantics></math> cells/day, (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>1500</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>2.41</mn> </mrow> </semantics></math> cells/day, panel (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>1.19</mn> </mrow> </semantics></math> cells/day, and (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>d</mi> <mo>=</mo> <mn>2.41</mn> </mrow> </semantics></math> cells/day. All illustrated trajectories relate to the following parameter values: <math display="inline"><semantics> <mrow> <mi>λ</mi> <mo>=</mo> <mn>1.865</mn> <mo>·</mo> <msup> <mn>10</mn> <mn>10</mn> </msup> </mrow> </semantics></math> cells, <math display="inline"><semantics> <mrow> <mi>ε</mi> <mo>=</mo> <mn>77,293</mn> </mrow> </semantics></math> cells, and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.95</mn> </mrow> </semantics></math>.</p>
Full article ">
26 pages, 4748 KiB  
Article
Reliable Energy Optimization Strategy for Fuel Cell Hybrid Electric Vehicles Considering Fuel Cell and Battery Health
by Cong Ji, Elkhatib Kamal and Reza Ghorbani
Energies 2024, 17(18), 4686; https://doi.org/10.3390/en17184686 - 20 Sep 2024
Cited by 1 | Viewed by 1212
Abstract
To enhance the fuel efficiency of fuel cell hybrid electric vehicles (FCHEVs), we propose a hierarchical energy management strategy (HEMS) to efficiently allocate power to a hybrid system comprising a fuel cell and a battery. Firstly, the upper-layer supervisor employs a fuzzy fault-tolerant [...] Read more.
To enhance the fuel efficiency of fuel cell hybrid electric vehicles (FCHEVs), we propose a hierarchical energy management strategy (HEMS) to efficiently allocate power to a hybrid system comprising a fuel cell and a battery. Firstly, the upper-layer supervisor employs a fuzzy fault-tolerant control and prediction strategy for the battery and fuel cell management system, ensuring vehicle stability and maintaining a healthy state of charge for both the battery and fuel cell, even during faults. Secondly, in the lower layer, dynamic programming and Pontryagin’s minimum principle are utilized to distribute the necessary power between the fuel cell system and the battery. This layer also incorporates an optimized proportional-integral controller for precise tracking of vehicle subsystem set-points. Finally, we compare the economic and dynamic performance of the vehicle using HEMS with other strategies, such as the equivalent consumption minimization strategy and fuzzy logic control strategy. Simulation results demonstrate that HEMS reduces hydrogen consumption and enhances overall vehicle energy efficiency across all operating conditions, indicating superior economic performance. Additionally, the dynamic performance of the vehicle shows significant improvement. Full article
Show Figures

Figure 1

Figure 1
<p>The main aims that can be considered for developing an EMS.</p>
Full article ">Figure 2
<p>Structure of the bus powertrain.</p>
Full article ">Figure 3
<p>Electrical model of the battery.</p>
Full article ">Figure 4
<p>The studied FCHEV modeled using TruckMaker software.</p>
Full article ">Figure 5
<p>The proposed overall control strategy.</p>
Full article ">Figure 6
<p>Schematic of the proposed FFTC.</p>
Full article ">Figure 7
<p>Block diagram for SOC prediction by adaptive fuzzy observer strategy.</p>
Full article ">Figure 8
<p>Block diagram for SOC prediction using ANFIS strategy.</p>
Full article ">Figure 9
<p>SOC prediction using ANFIS technique.</p>
Full article ">Figure 10
<p>Flowchart of DP algorithm.</p>
Full article ">Figure 11
<p>Algorithm of the rule-based EMS.</p>
Full article ">Figure 12
<p>Speed profile of the UDDS standard velocity profile.</p>
Full article ">Figure 13
<p>Power demand profile.</p>
Full article ">Figure 14
<p>Estimation of the SOC according to the proposed DP strategy with and without FTC.</p>
Full article ">Figure 15
<p>Evolution of <math display="inline"><semantics> <msub> <mi>H</mi> <mn>2</mn> </msub> </semantics></math> consumption according to the proposed DP strategy with and without FTC.</p>
Full article ">Figure 16
<p>Evolution of SOC according to the proposed strategies.</p>
Full article ">Figure 17
<p>Evolution of <math display="inline"><semantics> <msub> <mi>H</mi> <mn>2</mn> </msub> </semantics></math> consumption according to the proposed strategies.</p>
Full article ">
23 pages, 754 KiB  
Article
Modeling the Transmission Dynamics and Optimal Control Strategy for Huanglongbing
by Yujiang Liu, Shujing Gao, Di Chen and Bing Liu
Mathematics 2024, 12(17), 2648; https://doi.org/10.3390/math12172648 - 26 Aug 2024
Cited by 1 | Viewed by 727
Abstract
Huanglongbing (HLB), also known as citrus greening disease, represents a severe and imminent threat to the global citrus industry. With no complete cure currently available, effective control strategies are crucial. This article presents a transmission model of HLB, both with and without nutrient [...] Read more.
Huanglongbing (HLB), also known as citrus greening disease, represents a severe and imminent threat to the global citrus industry. With no complete cure currently available, effective control strategies are crucial. This article presents a transmission model of HLB, both with and without nutrient injection, to explore methods for controlling disease spread. By calculating the basic reproduction number (R0) and analyzing threshold dynamics, we demonstrate that the system remains globally stable when R0<1, but persists when R0>1. Sensitivity analyses reveal factors that significantly impact HLB spread on both global and local scales. We also propose a comprehensive optimal control model using the pontryagin minimum principle and validate its feasibility through numerical simulations. Results show that while removing infected trees and spraying insecticides can significantly reduce disease spread, a combination of measures, including the production of disease-free budwood and nursery trees, nutrient solution injection, removal of infected trees, and insecticide application, provides superior control and meets the desired control targets. These findings offer valuable insights for policymakers in understanding and managing HLB outbreaks. Full article
(This article belongs to the Section Mathematical Biology)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the formulated HLB disease dynamic model.</p>
Full article ">Figure 2
<p>Sensitivity analysis of the basic reproduction number, <math display="inline"><semantics> <msub> <mi mathvariant="script">R</mi> <mn>0</mn> </msub> </semantics></math>. The asterisk (*) denotes PRCCs that are significantly different from zero (<span class="html-italic">p</span>-value <math display="inline"><semantics> <mrow> <mo>&lt;</mo> <mn>0.01</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 3
<p>Comparison of control measures under strategy S1.</p>
Full article ">Figure 4
<p>Comparison of control measures under strategy S2.</p>
Full article ">Figure 5
<p>Comparison of control measures under Strategy S3.</p>
Full article ">Figure 6
<p>Comparison of control measures under Strategy S4.</p>
Full article ">Figure 7
<p>Control variables (<math display="inline"><semantics> <mrow> <msub> <mi>u</mi> <mn>1</mn> </msub> <mo>,</mo> <mspace width="3.33333pt"/> <msub> <mi>u</mi> <mn>2</mn> </msub> <mo>,</mo> <mspace width="3.33333pt"/> <msub> <mi>u</mi> <mn>3</mn> </msub> <mo>,</mo> <mspace width="3.33333pt"/> <msub> <mi>u</mi> <mn>4</mn> </msub> </mrow> </semantics></math>) of integrated control strategy (S5).</p>
Full article ">Figure 8
<p>Control variables (<math display="inline"><semantics> <mrow> <msub> <mi>u</mi> <mn>1</mn> </msub> <mo>,</mo> <mspace width="3.33333pt"/> <msub> <mi>u</mi> <mn>2</mn> </msub> <mo>,</mo> <mspace width="3.33333pt"/> <msub> <mi>u</mi> <mn>3</mn> </msub> <mo>,</mo> <mspace width="3.33333pt"/> <msub> <mi>u</mi> <mn>4</mn> </msub> </mrow> </semantics></math>) of integrated control strategy (S5).</p>
Full article ">
39 pages, 16376 KiB  
Article
Energy Management Systems’ Modeling and Optimization in Hybrid Electric Vehicles
by Yavuz Eray Altun and Osman Akın Kutlar
Energies 2024, 17(7), 1696; https://doi.org/10.3390/en17071696 - 2 Apr 2024
Cited by 2 | Viewed by 2766
Abstract
Optimization studies for the energy management systems of hybrid electric powertrains have critical importance as an effective measure for vehicle manufacturers to reduce greenhouse gas emissions and fuel consumption due to increasingly stringent emission regulations in the automotive industry, strict fuel economy legislation, [...] Read more.
Optimization studies for the energy management systems of hybrid electric powertrains have critical importance as an effective measure for vehicle manufacturers to reduce greenhouse gas emissions and fuel consumption due to increasingly stringent emission regulations in the automotive industry, strict fuel economy legislation, continuously rising oil prices, and increasing consumer awareness of global warming and environmental pollution. In this study, firstly, the mathematical model of the powertrain and the rule-based energy management system of the vehicle with a power-split hybrid electric vehicle configuration are developed in the Matlab/Simulink environment and verified with real test data from the vehicle dynamometer for the UDDS drive cycle. In this way, a realistic virtual test platform has been developed where the simulation results of the energy management systems based on discrete dynamic programming and Pontryagin’s minimum principle optimization can be used to train the artificial neural network-based energy management algorithms for hybrid electric vehicles. The average fuel consumption in relation to the break specific fuel consumption of the internal combustion engine and the total electrical energy consumption of the battery in relation to the operating efficiency of the electrical machines, obtained by comparing the simulation results at the initial battery charging conditions of the vehicle using different driving cycles, will be analyzed and the advantages of the different energy management techniques used will be evaluated. Full article
(This article belongs to the Special Issue Energy Management Control of Hybrid Electric Vehicles)
Show Figures

Figure 1

Figure 1
<p>Power-split hybrid electric vehicle configuration.</p>
Full article ">Figure 2
<p>Vehicle road load curve and battery open cycle voltage—cell internal resistance.</p>
Full article ">Figure 3
<p>Electric motor/generator MGA and MGB efficiencies and internal combustion engine BSFC.</p>
Full article ">Figure 4
<p>Hybrid electric vehicle driving modes.</p>
Full article ">Figure 5
<p>Hybrid electric vehicle Simulink models.</p>
Full article ">Figure 6
<p>Planetary gear ratio model.</p>
Full article ">Figure 7
<p>Battery open loop voltage and cell internal resistance [<a href="#B43-energies-17-01696" class="html-bibr">43</a>].</p>
Full article ">Figure 8
<p>Forces acting on the vehicle [<a href="#B21-energies-17-01696" class="html-bibr">21</a>].</p>
Full article ">Figure 9
<p>Drive cycles used for simulations.</p>
Full article ">Figure 10
<p>Battery and MGA and MGB validation results for UDDS-CS.</p>
Full article ">Figure 11
<p>Vehicle and transmission and ICE validation results for UDDS cycle.</p>
Full article ">Figure 12
<p>Hybrid mode selection diagram of HCU.</p>
Full article ">Figure 13
<p>Mode transition maps; engine load variation based on <span class="html-italic">SoC</span>.</p>
Full article ">Figure 14
<p>General structure of feedback neural networks [<a href="#B95-energies-17-01696" class="html-bibr">95</a>].</p>
Full article ">Figure 15
<p>ANN structure development via MATLAB.</p>
Full article ">Figure 16
<p>ANN mean square error (MSE) results with DDP and PMP CD/CS simulation data.</p>
Full article ">Figure 17
<p>ANN training regression results with DDP and PMP CD/CS simulation data.</p>
Full article ">Figure 18
<p>FTP-75 driving cycle CS mode simulation results.</p>
Full article ">Figure 19
<p>FTP75 driving cycle CD mode simulation results.</p>
Full article ">Figure 20
<p>HWFET driving cycle CS mode simulation results.</p>
Full article ">Figure 21
<p>HWFET driving cycle CD mode simulation results.</p>
Full article ">Figure 22
<p>JC08 driving cycle CS mode simulation results.</p>
Full article ">Figure 23
<p>JC08 driving cycle CD mode simulation results.</p>
Full article ">Figure 24
<p>NEDC driving cycle CS mode simulation results.</p>
Full article ">Figure 25
<p>NEDC driving cycle CD mode simulation results.</p>
Full article ">Figure 26
<p>UDDS driving cycle CS mode simulation results.</p>
Full article ">Figure 27
<p>UDDS driving cycle CD mode simulation results.</p>
Full article ">Figure 28
<p>WLTC driving cycle CS mode simulation results.</p>
Full article ">Figure 29
<p>WLTC driving cycle CD mode simulation results.</p>
Full article ">
23 pages, 916 KiB  
Article
Learning Fuel-Optimal Trajectories for Space Applications via Pontryagin Neural Networks
by Andrea D’Ambrosio and Roberto Furfaro
Aerospace 2024, 11(3), 228; https://doi.org/10.3390/aerospace11030228 - 14 Mar 2024
Cited by 4 | Viewed by 1399
Abstract
This paper demonstrates the utilization of Pontryagin Neural Networks (PoNNs) to acquire control strategies for achieving fuel-optimal trajectories. PoNNs, a subtype of Physics-Informed Neural Networks (PINNs), are tailored for solving optimal control problems through indirect methods. Specifically, PoNNs learn to solve the Two-Point [...] Read more.
This paper demonstrates the utilization of Pontryagin Neural Networks (PoNNs) to acquire control strategies for achieving fuel-optimal trajectories. PoNNs, a subtype of Physics-Informed Neural Networks (PINNs), are tailored for solving optimal control problems through indirect methods. Specifically, PoNNs learn to solve the Two-Point Boundary Value Problem derived from the application of the Pontryagin Minimum Principle to the problem’s Hamiltonian. Within PoNNs, the Extreme Theory of Functional Connections (X-TFC) is leveraged to approximate states and costates using constrained expressions (CEs). These CEs comprise a free function, modeled by a shallow neural network trained via Extreme Learning Machine, and a functional component that consistently satisfies boundary conditions analytically. Addressing discontinuous control, a smoothing technique is employed, substituting the sign function with a hyperbolic tangent function and implementing a continuation procedure on the smoothing parameter. The proposed methodology is applied to scenarios involving fuel-optimal Earth−Mars interplanetary transfers and Mars landing trajectories. Remarkably, PoNNs exhibit convergence to solutions even with randomly initialized parameters, determining the number and timing of control switches without prior information. Additionally, an analytical approximation of the solution allows for optimal control computation at unencountered points during training. Comparative analysis reveals the efficacy of the proposed approach, which rivals state-of-the-art methods such as the shooting technique and the adaptive Gaussian quadrature collocation method. Full article
(This article belongs to the Special Issue GNC for the Moon, Mars, and Beyond)
Show Figures

Figure 1

Figure 1
<p>Schematic summarizing how PoNNs work for learning the solution of generic OCPs. The red font in step 7 indicates the optimization procedure to retrieve the optimal weights <math display="inline"><semantics> <msubsup> <mi mathvariant="bold-italic">β</mi> <mi>j</mi> <mo>*</mo> </msubsup> </semantics></math>.</p>
Full article ">Figure 2
<p>Solution for the Earth−Mars interplanetary transfer without solution refinement.</p>
Full article ">Figure 3
<p>Earth−Mars interplanetary transfer trajectory after solution refinement. The purple arrows indicates the direction of the thrust.</p>
Full article ">Figure 4
<p>Thrust direction for the Earth−Mars interplanetary transfer after solution refinement.</p>
Full article ">Figure 5
<p>Solution for the Earth−Mars interplanetary transfer after solution refinement.</p>
Full article ">Figure 6
<p>Solution for the the Mars landing before the solution refinement.</p>
Full article ">Figure 7
<p>Mars landing trajectory after solution refinement.</p>
Full article ">Figure 8
<p>Thrust direction for the Mars landing after solution refinement.</p>
Full article ">Figure 9
<p>Solution for the Mars landing after solution refinement.</p>
Full article ">
17 pages, 9228 KiB  
Article
Research on Energy Distribution Strategy of Tandem Hybrid Tractor Based on the Pontryagin Minimum Principle
by Rundong Zhou, Lin Wang, Xiaoting Deng, Chao Su, Song Fang and Zhixiong Lu
Agriculture 2024, 14(3), 440; https://doi.org/10.3390/agriculture14030440 - 7 Mar 2024
Viewed by 1018
Abstract
In order to study the energy distribution of the tandem hybrid tractor and achieve optimal fuel economy under the whole operating condition, an energy distribution strategy based on PMP (Pontryagin minimum principle) is proposed. The performance parameters of the relevant power components are [...] Read more.
In order to study the energy distribution of the tandem hybrid tractor and achieve optimal fuel economy under the whole operating condition, an energy distribution strategy based on PMP (Pontryagin minimum principle) is proposed. The performance parameters of the relevant power components are obtained by constructing the test bench of the tandem hybrid tractor. At the same time, the mathematical model of energy distribution is established from the aspects of energy distribution objective function, state variable, and variable constraint. Based on the external characteristic test of the hub motor and the resistance analysis, the transport operation condition of the tractor is selected as the simulation target condition. The ADVISOR2002 and Simulink software are used to jointly simulate the three energy distribution strategies: the thermostat type, the power-following type, and the PMP type. The simulation results show that, compared with the other two, the fuel economy and battery power loss of the PMP-based energy distribution strategy are significantly improved. The fuel consumption per 100 km is decreased by 32.91% and 26.10%, respectively, which verifies the feasibility of the strategy. This study is of great significance for improving the production efficiency and reducing fuel consumption of hybrid tractors. Full article
(This article belongs to the Special Issue New Energy-Powered Agricultural Machinery and Equipment)
Show Figures

Figure 1

Figure 1
<p>Dynamic model of tandem hybrid tractor.</p>
Full article ">Figure 2
<p>Tandem hybrid tractor test bench.</p>
Full article ">Figure 3
<p>The curve of the external characteristic of the hub motor.</p>
Full article ">Figure 4
<p>Optimal working curve of engine/generator set.</p>
Full article ">Figure 5
<p>Relationship between terminal voltage and SOC in charging and discharging of power battery.</p>
Full article ">Figure 6
<p>The simplified model of power battery.</p>
Full article ">Figure 7
<p>Calculation flow of PMP global optimal control.</p>
Full article ">Figure 8
<p>Theoretical travel speed characteristics under transport condition.</p>
Full article ">Figure 9
<p>Simulation process of ADVISOR for tandem hybrid tractor.</p>
Full article ">Figure 10
<p>ADVISOR simulation model of tandem hybrid tractor.</p>
Full article ">Figure 11
<p>Tractor transport operating condition.</p>
Full article ">Figure 12
<p>Output power curve of engine/generator set.</p>
Full article ">Figure 13
<p>SOC curve of power battery.</p>
Full article ">Figure 14
<p>Loss of battery power.</p>
Full article ">Figure 15
<p>Output power curve of engine/generator set under PMP strategy.</p>
Full article ">Figure 16
<p>SOC curve of power battery under the PMP strategy.</p>
Full article ">Figure 17
<p>Influence of <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">λ</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mo>(</mo> <mi mathvariant="normal">S</mi> <mi mathvariant="normal">O</mi> <mi mathvariant="normal">C</mi> <mo>)</mo> </mrow> </semantics></math> under transportation condition.</p>
Full article ">Figure 18
<p>Loss of battery power.</p>
Full article ">
18 pages, 2601 KiB  
Article
Optimal Constrained Control of Arrays of Wave Energy Converters
by Habeebullah Abdulkadir and Ossama Abdelkhalik
J. Mar. Sci. Eng. 2024, 12(1), 104; https://doi.org/10.3390/jmse12010104 - 5 Jan 2024
Cited by 4 | Viewed by 1194
Abstract
Wave Energy Converters (WECs) are designed to be deployed in arrays, usually in a limited space, to minimize the cost of installation, mooring, and maintenance. Control methods that attempt to maximize the harvested power often lead to power flow from the WEC to [...] Read more.
Wave Energy Converters (WECs) are designed to be deployed in arrays, usually in a limited space, to minimize the cost of installation, mooring, and maintenance. Control methods that attempt to maximize the harvested power often lead to power flow from the WEC to the ocean, at times, to maximize the overall harvested power from the ocean over a longer period. The Power Take-Off (PTO) units that can provide power to the ocean (reactive power) are usually more expensive and complex. In this work, an optimal control formulation is presented using Pontryagin’s minimum principle that aims to maximize the harvested energy subject to constraints on the maximum PTO force and power flow direction. An analytical formulation is presented for the optimal control of an array of WECs, assuming irregular wave input. Three variations of the developed control are tested: a formulation without power constraints, a formulation that only allows for positive power, and finally, a formulation that allows for finite reactive power. The control is compared with optimally tuned damping and bang–bang control. Full article
(This article belongs to the Special Issue Tidal and Wave Energy)
Show Figures

Figure 1

Figure 1
<p>Layout of the devices in the array.</p>
Full article ">Figure 2
<p>Wave spectrum for the considered sea state.</p>
Full article ">Figure 3
<p>Excitation coefficients of the 3 devices.</p>
Full article ">Figure 4
<p>Added mass coefficients of the 3 devices.</p>
Full article ">Figure 5
<p>Radiation coefficients of the 3 devices.</p>
Full article ">Figure 6
<p>Control force of the power take-off units and power extracted by individual devices in the array.</p>
Full article ">Figure 7
<p>Total energy extracted by all the devices in the array.</p>
Full article ">Figure 8
<p>Cumulative power extracted by all the devices in the array.</p>
Full article ">Figure 9
<p>Displacement of all buoys when being controlled using damping control, bang–bang control, and the power-constrained bang-singular-bang control.</p>
Full article ">Figure 10
<p>Velocity of all buoys when being controlled using damping control, bang–bang control, and the power-constrained bang-singular-bang control.</p>
Full article ">Figure 11
<p>Power extracted by all devices when being controlled using damping control, bang–bang control, and the power-constrained bang-singular-bang control.</p>
Full article ">Figure 12
<p>Control force trajectory of the PTO values when being controlled using damping control, bang–bang control, and the power-constrained bang-singular-bang control.</p>
Full article ">Figure 13
<p>Total energy extracted from the waves when the devices are being controlled using damping control, bang–bang control, and the power-constrained bang-singular-bang control.</p>
Full article ">Figure 14
<p>Power extracted using PTO units when being controlled using damping control, bang–bang control, and the power-constrained bang-singular-bang control.</p>
Full article ">Figure 15
<p>Control force trajectories when being controlled using damping control, bang–bang control, and the power-constrained bang-singular-bang control.</p>
Full article ">Figure 16
<p>Total energy extracted by the devices when being controlled using damping control, bang–bang control, and the power-constrained bang-singular-bang control with <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>ϵ</mi> <mo>→</mo> </mover> <mo>=</mo> <mrow> <mn>0.5</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>4</mn> </msup> </mrow> <mspace width="4pt"/> <mi mathvariant="normal">W</mi> </mrow> </semantics></math>.</p>
Full article ">
15 pages, 8490 KiB  
Article
Optimal Control of Brushless Doubly Fed Wind Power Generator under Zero-Voltage Ride-Through
by Junyang Xu and Pengcheng Nie
Energies 2024, 17(1), 235; https://doi.org/10.3390/en17010235 - 1 Jan 2024
Cited by 2 | Viewed by 1170
Abstract
In the grid-connected operation dynamics of brushless doubly fed generators (BDFGs), a dip in the grid voltage is equivalent to suddenly adding a reverse voltage source at the parallel node. By deriving the expressions of the transient current of power winding (PW), control [...] Read more.
In the grid-connected operation dynamics of brushless doubly fed generators (BDFGs), a dip in the grid voltage is equivalent to suddenly adding a reverse voltage source at the parallel node. By deriving the expressions of the transient current of power winding (PW), control winding (CW), and rotor winding (RW) of a BDFG in the complex frequency domain under a natural state, it was concluded that the overshoot and oscillation time are affected by the CW voltage, the drop degree and phase of the grid voltage, and the rotor speed. Therefore, an optimal control strategy is proposed. A state model with the CW current as the state variable was constructed using the Pontryagin minimum principle. The finite-time integral value of the square of the electromagnetic torque was set as the objective function to achieve the minimum value that could suppress the overshoot and oscillation of the electromagnetic torque, and the optimal CW voltage command value was directly solved to accelerate the convergence of the BDFG’s physical quantities, thereby reducing the amplitude. Finally, the feasibility of the optimal control algorithm was verified using tests on an experimental platform. Full article
(This article belongs to the Special Issue Optimal Control of Wind and Wave Energy Converters)
Show Figures

Figure 1

Figure 1
<p>Equivalent schematic diagram of BDFG.</p>
Full article ">Figure 2
<p>Vector diagram of the rotating coordinate system.</p>
Full article ">Figure 3
<p>Steady-state equivalent circuit diagram of BDFG.</p>
Full article ">Figure 4
<p>Grid fault equivalent circuit diagram of BDFG.</p>
Full article ">Figure 5
<p>Short inductances of PW, CW, and RW. (<b>a</b>) PW short inductance, (<b>b</b>) CW short inductance, and (<b>c</b>) RW short inductance.</p>
Full article ">Figure 6
<p>Optimal control structure of the BDFG.</p>
Full article ">Figure 7
<p>Experimental platform of a brushless doubly fed power generation system.</p>
Full article ">Figure 8
<p>(<b>a</b>) Grid voltage, (<b>b</b>) PW current, (<b>c</b>) CW three-phase current, (<b>d</b>) PW and CW current dq components, (<b>e</b>) rotor speed, (<b>f</b>) electromagnetic torque, and (<b>g</b>) active and reactive power under PI control.</p>
Full article ">Figure 9
<p>(<b>a</b>) Grid voltage, (<b>b</b>) PW current, (<b>c</b>) CW three-phase current, (<b>d</b>) PW and CW current dq components, (<b>e</b>) rotor speed, (<b>f</b>) electromagnetic torque, and (<b>g</b>) active and reactive power under optimum control.</p>
Full article ">Figure 9 Cont.
<p>(<b>a</b>) Grid voltage, (<b>b</b>) PW current, (<b>c</b>) CW three-phase current, (<b>d</b>) PW and CW current dq components, (<b>e</b>) rotor speed, (<b>f</b>) electromagnetic torque, and (<b>g</b>) active and reactive power under optimum control.</p>
Full article ">Figure 10
<p>(<b>a</b>) Grid voltage, (<b>b</b>) PW current, (<b>c</b>) CW current, (<b>d</b>) PW and CW current dq components, (<b>e</b>) rotor speed, (<b>f</b>) electromagnetic torque, and (<b>g</b>) active and reactive power under flux linkage tracking.</p>
Full article ">
20 pages, 6645 KiB  
Article
An Energy Flow Control Algorithm of Regenerative Braking for Trams Based on Pontryagin’s Minimum Principle
by Ivan Župan, Viktor Šunde, Željko Ban and Branimir Novoselnik
Energies 2023, 16(21), 7346; https://doi.org/10.3390/en16217346 - 30 Oct 2023
Cited by 1 | Viewed by 1007
Abstract
Energy savings in electric rail transport are important in order to increase energy efficiency and reduce its carbon footprint. This can be achieved by storing and using the energy generated during regenerative braking. The system described in this paper consists of a supercapacitor [...] Read more.
Energy savings in electric rail transport are important in order to increase energy efficiency and reduce its carbon footprint. This can be achieved by storing and using the energy generated during regenerative braking. The system described in this paper consists of a supercapacitor energy storage system (SC ESS), a bidirectional DC/DC converter, and an algorithm to control the energy flow. The proper design of the algorithm is critical for maximizing energy savings and stabilizing the power grid, and it affects the lifetime of the SC ESS. This paper presents an energy flow control algorithm based on Pontryagin’s minimum principle that balances maximum energy savings with maximum SC ESS lifetime. The algorithm also performs SC ESS recharging while the rail vehicle stops on inclines to reduce the impact of its next acceleration on the power grid. To validate the algorithm, offline simulations are performed using real tram speed measurements. The results are then verified with a real-time laboratory emulation setup with HIL simulation. The tram and power grid are emulated with LiFePO4 batteries, while the SC ESS is emulated with a supercapacitor. The proposed algorithm controls a three-phase converter that enables energy exchange between the batteries and the supercapacitor. The results show that the proposed algorithm is feasible in real time and that it can be used under real operating conditions. Full article
(This article belongs to the Special Issue Advances in Energy Storage Systems for Renewable Energy)
Show Figures

Figure 1

Figure 1
<p>Regenerative braking system model with included supercapacitor electrothermal model.</p>
Full article ">Figure 2
<p>Simulation model of the block for estimating the inclination of the track.</p>
Full article ">Figure 3
<p>Track elevation for line No. 14 determined using the Google Earth application (blue) and the estimated inclination of the track for line No. 14 (orange).</p>
Full article ">Figure 4
<p>SC thermal model.</p>
Full article ">Figure 5
<p>Energy flows in a simplified model of the regenerative braking system.</p>
Full article ">Figure 6
<p>Simulation model of the supercapacitor, power grid, and tram system, together with the control algorithm.</p>
Full article ">Figure 7
<p>Tram speed profile on line No. 14 of the Zagreb tram network.</p>
Full article ">Figure 8
<p>Power grid voltage (<b>a</b>) during the entire ride, (<b>b</b>) detail during the ascent (arrows show the moments of recharging the SC).</p>
Full article ">Figure 9
<p>SC current (<b>a</b>) during the entire ride, (<b>b</b>) detail during the ascent (arrows show the moments of recharging the SC).</p>
Full article ">Figure 10
<p>SC temperature waveform.</p>
Full article ">Figure 11
<p>Structure of the HIL laboratory setup.</p>
Full article ">Figure 12
<p>HIL simulation model within the Typhoon HIL programming interface.</p>
Full article ">Figure 13
<p>Laboratory setup for HIL simulation experiment: 1—autotransformer, 2—Danfoss FC302 converter, 3—Typhoon HIL402, 4, 5—LiFePO4 batteries, 6—Maxwell BMOD0083 P048 B01 supercapacitor, 7—fluke multimeter.</p>
Full article ">Figure 14
<p>Denormalized currents of the supercapacitor from the HIL simulation experiment compared with the results of the offline simulation for tram driving: (<b>a</b>) on a flat part of the track, (<b>b</b>) ascent of the track, (<b>c</b>) descent of the track.</p>
Full article ">Figure 14 Cont.
<p>Denormalized currents of the supercapacitor from the HIL simulation experiment compared with the results of the offline simulation for tram driving: (<b>a</b>) on a flat part of the track, (<b>b</b>) ascent of the track, (<b>c</b>) descent of the track.</p>
Full article ">
40 pages, 869 KiB  
Article
Energy-Efficient Train Driving Based on Optimal Control Theory
by Wolfram Heineken, Marc Richter and Torsten Birth-Reichert
Energies 2023, 16(18), 6712; https://doi.org/10.3390/en16186712 - 19 Sep 2023
Viewed by 1293
Abstract
Efficient train driving plays a vital role in reducing the overall energy consumption in the railway sector. An energy minimising control strategy can be computed using the framework given by optimal control theory; in particular, the Pontryagin maximum principle can be used. Our [...] Read more.
Efficient train driving plays a vital role in reducing the overall energy consumption in the railway sector. An energy minimising control strategy can be computed using the framework given by optimal control theory; in particular, the Pontryagin maximum principle can be used. Our optimisation approach is based on an algorithm presented by Khmelnitsky that considers electric trains equipped with regenerative braking. A derivation of Khmelnitsky’s theory from a more general formulation of the maximum principle is given in this article, and a complete list of switching cases between different driving regimes is included that is essential for practical application. A number of numerical examples are added to visualise the various switching cases. Energy consumption data from real-life operation of passenger trains are compared to the calculated energy minimum. In the presented study, the optimised strategy was able to save 37 percent of the average energy demand of the train in operation. The sensitivity of the energy consumption to deviations of the train speed from the optimum speed profile is studied in an example. Another example illustrates that the efficiency of regenerative braking has an effect on the optimum speed profile. Full article
(This article belongs to the Section E: Electric Vehicles)
Show Figures

Figure 1

Figure 1
<p>Maximum force for traction and regenerative brake for an electric locomotive with mass <math display="inline"><semantics><mrow><msub><mi>m</mi><mi>loc</mi></msub><mo>=</mo><mn>84</mn><mo> </mo><mi mathvariant="normal">t</mi></mrow></semantics></math> and <math display="inline"><semantics><mrow><msub><mi>P</mi><mrow><mi>tr</mi><mo>,</mo><mi>mech</mi></mrow></msub><mo>=</mo><msub><mi>P</mi><mrow><mi>br</mi><mo>,</mo><mi>mech</mi></mrow></msub><mo>=</mo><mn>5.6</mn><mrow><mo> </mo><mi>MW</mi></mrow></mrow></semantics></math>.</p>
Full article ">Figure 2
<p>Track altitude in Example 1.</p>
Full article ">Figure 3
<p>Example 1: pcs-intervals PT and PB, fastest motion, and numbering.</p>
Full article ">Figure 4
<p>Example 1: <span class="html-italic">K</span>-trajectories starting from <math display="inline"><semantics><mrow><mi>s</mi><mo>=</mo><mn>0</mn></mrow></semantics></math> (port 1) for various values of <math display="inline"><semantics><mi>ψ</mi></semantics></math> at <math display="inline"><semantics><mrow><mi>s</mi><mo>=</mo><mn>0</mn></mrow></semantics></math>. Trajectories change from full traction to coasting and then to full regenerative braking. The blue trajectory connects port 1 with port 2.</p>
Full article ">Figure 5
<p>Example 1: <math display="inline"><semantics><mi>ψ</mi></semantics></math>-trajectories starting from <math display="inline"><semantics><mrow><mi>s</mi><mo>=</mo><mn>0</mn></mrow></semantics></math> (port 1) for various values of <math display="inline"><semantics><mi>ψ</mi></semantics></math> at <math display="inline"><semantics><mrow><mi>s</mi><mo>=</mo><mn>0</mn></mrow></semantics></math>. Trajectories change from full traction to coasting and then to full regenerative braking. The blue trajectory connects port 1 with port 2.</p>
Full article ">Figure 6
<p>Example 1: <span class="html-italic">K</span>-trajectories starting from PT-interval with number 3.</p>
Full article ">Figure 7
<p>Example 1: <math display="inline"><semantics><mi>ψ</mi></semantics></math>-trajectories starting from PT-interval with number 3.</p>
Full article ">Figure 8
<p>Example 1: <span class="html-italic">K</span>-trajectories starting from interval 4, heading for interval 8.</p>
Full article ">Figure 9
<p>Example 1: <math display="inline"><semantics><mi>ψ</mi></semantics></math>-trajectories starting from interval 4, heading for interval 8.</p>
Full article ">Figure 10
<p>Example 2: <span class="html-italic">K</span>-trajectories.</p>
Full article ">Figure 11
<p>Example 2: <math display="inline"><semantics><mi>ψ</mi></semantics></math>-trajectories.</p>
Full article ">Figure 12
<p>Example 2: Zoom into <math display="inline"><semantics><mi>ψ</mi></semantics></math>-trajectories.</p>
Full article ">Figure 13
<p>Example 3: <span class="html-italic">K</span>-trajectories.</p>
Full article ">Figure 14
<p>Example 3: <math display="inline"><semantics><mi>ψ</mi></semantics></math>-trajectories.</p>
Full article ">Figure 15
<p>Example 4: <span class="html-italic">K</span>-trajectories.</p>
Full article ">Figure 16
<p>Example 4: <math display="inline"><semantics><mi>ψ</mi></semantics></math>-trajectories.</p>
Full article ">Figure 17
<p>Example 4: train speed.</p>
Full article ">Figure 18
<p>Example 4: net electric energy.</p>
Full article ">Figure 19
<p>Comparison of energy supply <math display="inline"><semantics><mrow><msub><mi>E</mi><mi>tr</mi></msub><mo>+</mo><msub><mi>E</mi><mi>add</mi></msub></mrow></semantics></math>; measurement vs. model.</p>
Full article ">Figure 20
<p>Total energy demand <math display="inline"><semantics><msub><mi>E</mi><mi>tot</mi></msub></semantics></math> over reference time <span class="html-italic">T</span> for track in <a href="#sec1-energies-16-06712" class="html-sec">Section 1</a>.</p>
Full article ">Figure 21
<p>Total energy demand <math display="inline"><semantics><msub><mi>E</mi><mi>tot</mi></msub></semantics></math> over reference time <span class="html-italic">T</span> for track <a href="#sec2-energies-16-06712" class="html-sec">Section 2</a>.</p>
Full article ">Figure 22
<p>Total energy demand <math display="inline"><semantics><msub><mi>E</mi><mi>tot</mi></msub></semantics></math> over reference time <span class="html-italic">T</span> for track in <a href="#sec3-energies-16-06712" class="html-sec">Section 3</a>.</p>
Full article ">Figure 23
<p>Total energy demand <math display="inline"><semantics><msub><mi>E</mi><mi>tot</mi></msub></semantics></math> over reference time <span class="html-italic">T</span> for track in <a href="#sec4-energies-16-06712" class="html-sec">Section 4</a>.</p>
Full article ">Figure 24
<p>Energy ratio <math display="inline"><semantics><msub><mi>η</mi><mi>i</mi></msub></semantics></math> for each run, shown in descending order.</p>
Full article ">Figure 25
<p>Example 5: track altitude <span class="html-italic">z</span>.</p>
Full article ">Figure 26
<p>Example 5: train speed <span class="html-italic">v</span> (optimum and variants 1–3).</p>
Full article ">Figure 27
<p>Example 5: total electric energy <math display="inline"><semantics><msub><mi>E</mi><mi>tot</mi></msub></semantics></math> (optimum and variants 1–3).</p>
Full article ">Figure 28
<p>Example 5: total electric energy <math display="inline"><semantics><msub><mi>E</mi><mi>tot</mi></msub></semantics></math> (optimum and variants 1–3). Zoom into <a href="#energies-16-06712-f027" class="html-fig">Figure 27</a>.</p>
Full article ">Figure 29
<p>Example 5: train speed <span class="html-italic">v</span> (optimum and variants 4–7).</p>
Full article ">Figure 30
<p>Example 5: total electric energy <math display="inline"><semantics><msub><mi>E</mi><mi>tot</mi></msub></semantics></math> (optimum and variants 4–7).</p>
Full article ">Figure 31
<p>Example 5: total electric energy <math display="inline"><semantics><msub><mi>E</mi><mi>tot</mi></msub></semantics></math> (optimum and variants 4–7). Zoom into <a href="#energies-16-06712-f030" class="html-fig">Figure 30</a>.</p>
Full article ">Figure 32
<p>Example 6: train speed <span class="html-italic">v</span> for different regenerative brake efficiency <math display="inline"><semantics><msub><mi>η</mi><mi>br</mi></msub></semantics></math>.</p>
Full article ">Figure 33
<p>Example 6: total electric energy <math display="inline"><semantics><msub><mi>E</mi><mi>tot</mi></msub></semantics></math> for different regenerative brake efficiency <math display="inline"><semantics><msub><mi>η</mi><mi>br</mi></msub></semantics></math>.</p>
Full article ">
31 pages, 1755 KiB  
Article
Time-Optimal Problem in the Roto-Translation Group with Admissible Control in a Circular Sector
by Alexey Mashtakov and Yuri Sachkov
Mathematics 2023, 11(18), 3931; https://doi.org/10.3390/math11183931 - 15 Sep 2023
Cited by 1 | Viewed by 898
Abstract
We study a time-optimal problem in the roto-translation group with admissible control in a circular sector. The problem reveals the trajectories of a car model that can move forward on a plane and turn with a given minimum turning radius. Our work generalizes [...] Read more.
We study a time-optimal problem in the roto-translation group with admissible control in a circular sector. The problem reveals the trajectories of a car model that can move forward on a plane and turn with a given minimum turning radius. Our work generalizes the sub-Riemannian problem by adding a restriction on the velocity vector to lie in a circular sector. The sub-Riemannian problem is given by a special case when the sector is the full disc. The trajectories of the system are applicable in image processing to detect salient lines. We study the local and global controllability of the system and the existence of a solution for given arbitrary boundary conditions. In a general case of the sector opening angle, the system is globally but not small-time locally controllable. We show that when the angle is obtuse, a solution exists for any boundary conditions, and when the angle is reflex, a solution does not exist for some boundary conditions. We apply the Pontryagin maximum principle and derive a Hamiltonian system for extremals. Analyzing a phase portrait of the Hamiltonian system, we introduce the rectified coordinates and obtain an explicit expression for the extremals in Jacobi elliptic functions. We show that abnormal extremals are of circular type, and they correspond to motions of a car along circular arcs of minimal possible radius. The normal extremals in a general case are given by concatenation of segments of sub-Riemannian geodesics in SE2 and arcs of circular extremals. We show that, in a general case, the vertical (momentum) part of the extremals is periodic. We partially study the optimality of the extremals and provide estimates for the cut time in terms of the period of the vertical part. Full article
(This article belongs to the Special Issue Variational Methods on Riemannian Manifolds: Theory and Applications)
Show Figures

Figure 1

Figure 1
<p>A model of a car that can move forward and turn within a given minimal radius. The control <math display="inline"><semantics> <msub> <mi>u</mi> <mn>1</mn> </msub> </semantics></math> is responsible for moving forward and <math display="inline"><semantics> <msub> <mi>u</mi> <mn>2</mn> </msub> </semantics></math> for the turn.</p>
Full article ">Figure 2
<p>Set of admissible controls for various models of a car on a plane: Dubins car [<a href="#B1-mathematics-11-03931" class="html-bibr">1</a>]; Reeds–Shepp car [<a href="#B2-mathematics-11-03931" class="html-bibr">2</a>]; Ardentov model [<a href="#B3-mathematics-11-03931" class="html-bibr">3</a>] (generalized Dubins car); Sachkov model [<a href="#B4-mathematics-11-03931" class="html-bibr">4</a>] (sub-Riemannian problem); Berestovskii model [<a href="#B5-mathematics-11-03931" class="html-bibr">5</a>]; Duits model [<a href="#B6-mathematics-11-03931" class="html-bibr">6</a>]; our model with control in a sector.</p>
Full article ">Figure 3
<p>The maximum condition.</p>
Full article ">Figure 4
<p>Abnormal case. (<b>Left</b>) Level surfaces of the Hamiltonian <math display="inline"><semantics> <mrow> <mi>H</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> (in green) and the Casimir <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math> (in red). (<b>Center</b>) Phase portrait on the surface <math display="inline"><semantics> <mrow> <mi>H</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>. (<b>Right</b>) An abnormal extremal trajectory.</p>
Full article ">Figure 5
<p>Level surfaces of the Hamiltonian <span class="html-italic">H</span> (in green) and the Casimir <span class="html-italic">E</span> (in red). (<b>Left</b>) <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>&lt;</mo> <mfrac> <mn>1</mn> <mn>2</mn> </mfrac> </mrow> </semantics></math>. (<b>Center</b>) <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>=</mo> <mfrac> <mn>1</mn> <mn>2</mn> </mfrac> </mrow> </semantics></math>. (<b>Right</b>) <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>&gt;</mo> <mfrac> <mn>1</mn> <mn>2</mn> </mfrac> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>The phase portrait on the level surfaces of the Hamiltonian <math display="inline"><semantics> <mrow> <mi>H</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Stratification of the domain of the vertical part.</p>
Full article ">Figure 8
<p>Timeline with indicated instances of switching and the corresponding trajectory.</p>
Full article ">Figure 9
<p>Rectified coordinates in the domain <math display="inline"><semantics> <mrow> <mo form="prefix">cos</mo> <mi>α</mi> <mo>&lt;</mo> <mi mathvariant="script">E</mi> <mo>&lt;</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>Timeline for the trajectory with indicated instances of switches.</p>
Full article ">Figure 11
<p>Two extremal trajectories in <math display="inline"><semantics> <mrow> <msubsup> <mi>S</mi> <mn>1</mn> <mo>±</mo> </msubsup> <mo>∪</mo> <msubsup> <mi>O</mi> <mn>1</mn> <mo>±</mo> </msubsup> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mfrac> <mrow> <mn>3</mn> <mi>π</mi> </mrow> <mn>7</mn> </mfrac> </mrow> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>15.4</mn> <mo>]</mo> </mrow> </semantics></math>: for <math display="inline"><semantics> <mrow> <msup> <mi>h</mi> <mn>0</mn> </msup> <mo>=</mo> <mrow> <mo>(</mo> <mn>0.32</mn> <mo>,</mo> <mo>−</mo> <mn>0.85</mn> <mo>,</mo> <mo>−</mo> <mn>0.66</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msup> <mi>h</mi> <mn>0</mn> </msup> <mo>=</mo> <mrow> <mo>(</mo> <mn>0.52</mn> <mo>,</mo> <mn>0.85</mn> <mo>,</mo> <mo>−</mo> <mn>0.46</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>. The sub-Riemannian arcs are depicted in red, and arcs of the circles are depicted in blue.</p>
Full article ">Figure 12
<p>Rectified coordinates in the domain <math display="inline"><semantics> <mrow> <mi mathvariant="script">E</mi> <mo>&gt;</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 13
<p>Two extremal trajectories in <math display="inline"><semantics> <mrow> <msubsup> <mi>S</mi> <mn>2</mn> <mo>±</mo> </msubsup> <mo>∪</mo> <msubsup> <mi>O</mi> <mn>2</mn> <mo>±</mo> </msubsup> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mfrac> <mi>π</mi> <mn>4</mn> </mfrac> </mrow> </semantics></math>: for <math display="inline"><semantics> <mrow> <msup> <mi>h</mi> <mn>0</mn> </msup> <mo>=</mo> <mrow> <mo>(</mo> <mn>0.7</mn> <mo>,</mo> <mo>−</mo> <mn>0.714</mn> <mo>,</mo> <mo>−</mo> <mn>1.05</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msup> <mi>h</mi> <mn>0</mn> </msup> <mo>=</mo> <mrow> <mo>(</mo> <mn>0.7</mn> <mo>,</mo> <mn>0.714</mn> <mo>,</mo> <mo>−</mo> <mn>0.85</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>, for <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>∈</mo> <mo>[</mo> <mn>0</mn> <mo>,</mo> <mn>16</mn> <mo>]</mo> </mrow> </semantics></math>. The sub-Riemannian arcs are depicted in red, and arcs of the circles are depicted in blue.</p>
Full article ">Figure 14
<p>Rectified coordinates in the domain <math display="inline"><semantics> <mrow> <mi mathvariant="script">E</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>h</mi> <mn>1</mn> </msub> <mo>&lt;</mo> <mn>1</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 15
<p>Three extremal trajectories in <math display="inline"><semantics> <mrow> <msubsup> <mi>S</mi> <mn>3</mn> <mrow> <mo>±</mo> <mo>±</mo> </mrow> </msubsup> <mo>∪</mo> <msubsup> <mi>O</mi> <mn>3</mn> <mo>±</mo> </msubsup> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mfrac> <mi>π</mi> <mn>4</mn> </mfrac> </mrow> </semantics></math>: for <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msub> <mi>h</mi> <mn>10</mn> </msub> <mo>,</mo> <msub> <mi>s</mi> <mn>20</mn> </msub> <mo>,</mo> <msub> <mi>s</mi> <mn>30</mn> </msub> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>0.95</mn> <mo>,</mo> <mo>−</mo> <mn>1</mn> <mo>,</mo> <mn>1</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>12</mn> </mrow> </semantics></math>; for <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msub> <mi>h</mi> <mn>10</mn> </msub> <mo>,</mo> <msub> <mi>s</mi> <mn>20</mn> </msub> <mo>,</mo> <msub> <mi>s</mi> <mn>30</mn> </msub> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>0.99999</mn> <mo>,</mo> <mn>1</mn> <mo>,</mo> <mn>1</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>21</mn> </mrow> </semantics></math>; for <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <msub> <mi>h</mi> <mn>10</mn> </msub> <mo>,</mo> <msub> <mi>s</mi> <mn>20</mn> </msub> <mo>,</mo> <msub> <mi>s</mi> <mn>30</mn> </msub> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mn>0.9</mn> <mo>,</mo> <mn>1</mn> <mo>,</mo> <mn>1</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>15.5</mn> </mrow> </semantics></math>. The sub-Riemannian arcs are depicted in red, and arcs of the circles are depicted in blue.</p>
Full article ">Figure 16
<p>Extremal trajectories in <math display="inline"><semantics> <msubsup> <mi>O</mi> <mn>4</mn> <mo>±</mo> </msubsup> </semantics></math>.</p>
Full article ">Figure 17
<p>Extremal trajectory in <math display="inline"><semantics> <msub> <mi>S</mi> <mn>5</mn> </msub> </semantics></math>.</p>
Full article ">Figure 18
<p>Nonoptimal arc of the extremal trajectory.</p>
Full article ">
15 pages, 589 KiB  
Article
An Analytic Method to Determine the Optimal Time for the Induction Phase of Anesthesia
by Mohamed A. Zaitri, Cristiana J. Silva and Delfim F. M. Torres
Axioms 2023, 12(9), 867; https://doi.org/10.3390/axioms12090867 - 8 Sep 2023
Cited by 1 | Viewed by 1043
Abstract
We obtain an analytical solution for the time-optimal control problem in the induction phase of anesthesia. Our solution is shown to align numerically with the results obtained from the conventional shooting method. The induction phase of anesthesia relies on a pharmacokinetic/pharmacodynamic (PK/PD) model [...] Read more.
We obtain an analytical solution for the time-optimal control problem in the induction phase of anesthesia. Our solution is shown to align numerically with the results obtained from the conventional shooting method. The induction phase of anesthesia relies on a pharmacokinetic/pharmacodynamic (PK/PD) model proposed by Bailey and Haddad in 2005 to regulate the infusion of propofol. In order to evaluate our approach and compare it with existing results in the literature, we examine a minimum-time problem for anesthetizing a patient. By applying the Pontryagin minimum principle, we introduce the shooting method as a means to solve the problem at hand. Additionally, we conducted numerical simulations using the MATLAB computing environment. We solve the time-optimal control problem using our newly proposed analytical method and discover that the optimal continuous infusion rate of the anesthetic and the minimum required time for transition from the awake state to an anesthetized state exhibit similarity between the two methods. However, the advantage of our new analytic method lies in its independence from unknown initial conditions for the adjoint variables. Full article
(This article belongs to the Special Issue Mathematical Methods in the Applied Sciences)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the PK/PD model with the effect site compartment of Bailey and Haddad [<a href="#B19-axioms-12-00867" class="html-bibr">19</a>].</p>
Full article ">Figure 2
<p>The state trajectory, controlled BIS index, and trajectory of the fast states corresponding to the optimal control <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> of <a href="#axioms-12-00867-f004" class="html-fig">Figure 4</a>, using the shooting method.</p>
Full article ">Figure 3
<p>The state trajectory, controlled BIS index, and trajectory of the fast states corresponding to the optimal control <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> of <a href="#axioms-12-00867-f004" class="html-fig">Figure 4</a>, using the analytical method.</p>
Full article ">Figure 4
<p>The optimal continuous infusion rate <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> of the induction phase of anesthesia, as obtained by the shooting and analytical methods.</p>
Full article ">
14 pages, 1216 KiB  
Article
One New Property of a Class of Linear Time-Optimal Control Problems
by Borislav Penev
Mathematics 2023, 11(16), 3486; https://doi.org/10.3390/math11163486 - 11 Aug 2023
Viewed by 1385
Abstract
The following paper deals with a new property of linear time-optimal control problems with real eigenvalues of the system. This property unveils the possibility of synthesizing the time-optimal control without describing the switching hyper-surfaces. Furthermore, the novel technique offers an alternative solution to [...] Read more.
The following paper deals with a new property of linear time-optimal control problems with real eigenvalues of the system. This property unveils the possibility of synthesizing the time-optimal control without describing the switching hyper-surfaces. Furthermore, the novel technique offers an alternative solution to the classic example of the time-optimal control of a double integrator system. Full article
(This article belongs to the Special Issue Mathematical Modelling, Simulation, and Optimal Control)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the initial system (1) in form (10).</p>
Full article ">Figure 2
<p>Representation of the areas <math display="inline"><semantics> <mrow> <msup> <mi>R</mi> <mo>+</mo> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msup> <mi>R</mi> <mo>−</mo> </msup> </mrow> </semantics></math> as well as the two parts <math display="inline"><semantics> <mrow> <msup> <mi>γ</mi> <mo>+</mo> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msup> <mi>γ</mi> <mo>−</mo> </msup> </mrow> </semantics></math> of the switching curve <math display="inline"><semantics> <mrow> <msub> <mi>S</mi> <mn>2</mn> </msub> </mrow> </semantics></math> in the phase plane <math display="inline"><semantics> <mrow> <mi>y</mi> <mi>v</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Near time-optimal process with an accuracy of <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>0.001</mn> </mrow> </semantics></math> referring to the initial state: (<b>a</b>) <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <msub> <mi>y</mi> <mn>0</mn> </msub> <mo>,</mo> <msub> <mi>v</mi> <mn>0</mn> </msub> </mrow> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mrow> <mn>10</mn> <mo>,</mo> <mo> </mo> <mn>0</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> with corresponding <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <msub> <mi>x</mi> <mrow> <mn>10</mn> </mrow> </msub> <mo>,</mo> <msub> <mi>x</mi> <mrow> <mn>20</mn> </mrow> </msub> </mrow> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mrow> <mn>0</mn> <mo>,</mo> <mo> </mo> <mn>10</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <msub> <mi>y</mi> <mn>0</mn> </msub> <mo>,</mo> <msub> <mi>v</mi> <mn>0</mn> </msub> </mrow> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mrow> <mo>−</mo> <mn>10</mn> <mo>,</mo> <mo> </mo> <mn>0</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> with corresponding <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <msub> <mi>x</mi> <mrow> <mn>10</mn> </mrow> </msub> <mo>,</mo> <msub> <mi>x</mi> <mrow> <mn>20</mn> </mrow> </msub> </mrow> <mo>)</mo> </mrow> <mo>=</mo> <mrow> <mo>(</mo> <mrow> <mn>0</mn> <mo>,</mo> <mo>−</mo> <mn>10</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Phase trajectories of the near time-optimal processes with an accuracy of <math display="inline"><semantics> <mrow> <msub> <mi>ε</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>0.001</mn> </mrow> </semantics></math> in the phase plane: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>y</mi> <mi>v</mi> </mrow> </semantics></math> of the system (32); (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>1</mn> </msub> <msub> <mi>x</mi> <mn>2</mn> </msub> </mrow> </semantics></math> of the system (44).</p>
Full article ">
Back to TopTop