key: cord-0911471-t7e80tog authors: Wacker, Benjamin; Schlüter, Jan title: Time-continuous and time-discrete SIR models revisited: theory and applications date: 2020-10-07 journal: Adv Differ Equ DOI: 10.1186/s13662-020-02995-1 sha: ba63b8779463bbad1e439de500306bcddd83a0d4 doc_id: 911471 cord_uid: t7e80tog Since Kermack and McKendrick have introduced their famous epidemiological SIR model in 1927, mathematical epidemiology has grown as an interdisciplinary research discipline including knowledge from biology, computer science, or mathematics. Due to current threatening epidemics such as COVID-19, this interest is continuously rising. As our main goal, we establish an implicit time-discrete SIR (susceptible people–infectious people–recovered people) model. For this purpose, we first introduce its continuous variant with time-varying transmission and recovery rates and, as our first contribution, discuss thoroughly its properties. With respect to these results, we develop different possible time-discrete SIR models, we derive our implicit time-discrete SIR model in contrast to many other works which mainly investigate explicit time-discrete schemes and, as our main contribution, show unique solvability and further desirable properties compared to its continuous version. We thoroughly show that many of the desired properties of the time-continuous case are still valid in the time-discrete implicit case. Especially, we prove an upper error bound for our time-discrete implicit numerical scheme. Finally, we apply our proposed time-discrete SIR model to currently available data regarding the spread of COVID-19 in Germany and Iran. Since its outbreak in Wuhan (China) in December 2019, the quick spread of COVID-19 has threatened people worldwide. Politicians around the globe have to balance different interests and need to make tremendous decisions which impact our daily life. For these reasons, governments around the world heavily rely on scientific advice in the present situation. Thus, John Hopkins University has been collecting epidemiological data regarding COVID-19 from many countries during the last months [1, 2] . Additionally, many biological and medical studies regarding different aspects of this new coronavirus have been rapidly appearing in scientific journals [3] [4] [5] [6] [7] [8] [9] . However, to estimate the impact of COVID-19, governments need forecasts from as many models as possible. Kermack and McKendrick introduced the now well-known SIR model in one of mathematical epidemiology's most groundbreaking works in 1927 [10] . They assumed a fixed population size and divided this population into three different homogeneous groups of people, namely susceptible people, infectious people, and recovered people, excluding births, deaths, and deaths by disease from their model. Due to its success and simplicity, their work was reprinted in 1991 [11] [12] [13] . In upcoming decades, epidemiologists and mathematicians have developed many variants and extensions of this basic model by, for example, adding age or spatial structures [14] [15] [16] [17] [18] [19] [20] . After the outbreak of COVID-19, many scientists have recently published articles with emphasis on epidemic forecasts which strongly relate to mathematical models. Many approaches, mainly focusing on stochastic arguments, with respect to predicting forecasts of the total number of infected people have appeared during the last weeks [21] [22] [23] [24] [25] [26] or in the past [27, 28] . Recently, neural networks have been applied to forecasting [29] and, additionally, different authors such as Atangana, Baleanu, or Khan have used fractional differential equations in extended SIR-type models to investigate the spread of COVID-19 or mathematical biology in general [30] [31] [32] [33] [34] . There are works with respect to SIR models [35] [36] [37] and their time-discrete versions [38] . However, one finds mainly explicit schemes with respect to time-discrete SIR models in the aforementioned works and references therein. For this reason, we summarize and extend some results on properties of the time-continuous classical SIR model, and we propose an implicit time-discrete version of this classical SIR model and prove that this timediscrete variant maintains many of time-continuous version's properties. Hence, the aim of our study is to propose a nonautonomous SIR model, analyze the properties of its timecontinuous formulation, and develop an implicit numerical solution algorithm where the main properties of the time-continuous variant are conserved. More precisely, our main contributions can be summarized as follows: 1) At first, we propose a modified time-continuous SIR model with time-varying transmission and recovery rates; 2) Secondly, we conclude well-posedness of our time-continuous problem formulation. This includes global existence in time, global uniqueness in time, based on an inductive application of Banach's fixed point theorem, and continuous dependence on initial conditions and time-varying rates; 3) In case of the time-discrete implicit model, we provide unique solvability, monotonicity properties, and an upper error bound between the solution of the implicit time-discrete problem formulation and the solution of the time-continuous problem formulation; 4) Finally, we compare our model predictions with real-world data regarding the spread of COVID-19 from different countries. Data have been collected by John Hopkins University. Conclusively, we summarize our results and give a brief outlook on possible extensions. In this section, we portray the time-continuous SIR model and its properties. Here, we recall Lipschitz continuity of a function on Euclidean spaces. Definition 1 ([39, Sect. 3.2] ) Let d 1 , d 2 ∈ N. If S ⊂ R d 1 , a defined function F : S − → R d 2 is called Lipschitz continuous on S if there exists a nonnegative constant L ≥ 0 such that holds for all x, y ∈ S. Here, · denotes a suitable norm on the corresponding Euclidean space. We shall call F locally Lipschitz continuous if for every point x 0 ∈ U there exists a neighborhood V of x 0 such that the restriction of F to V is Lipschitz continuous on V . In a more general framework, we consider a nonlinear initial value problem ⎧ ⎨ ⎩ z (t) = G(t, z(t)), where we define our solution vector z(t) = (x 1 (t), . . . , x n (t)) T , our vectorial function G(t, z(t)) = (g 1 (t, z(t)), . . . , g n (t, z(t))) T , and our initial condition z 0 ∈ R n . To conclude global existence in time, we can apply the following theorem that is a direct consequence of Grönwall's lemma. holds for all z(t) ∈ R n , then the solution of the initial value problem (2) exists for all time t ∈ R and, moreover, it holds for all t ∈ R. To prove global uniqueness in time, we need Banach's fixed point theorem since fixed point theorems have been successfully applied as a versatile tool in different areas of mathematics [40, 41] . Theorem 2 ([42, Theorem V.18]) Let (X, ) be a complete metric space with the metric mapping : X × X − → [0, ∞). Let T : X − → X be a strict contraction, i.e., there exists a constant K ∈ [0, 1) such that (Tx, Ty) ≤ K · (x, y) holds for all x, y ∈ X. Then the mapping T has a unique fixed point. Finally, since we want to provide continuous dependence on initial conditions and other data, we need the following inequality named after Gronwall and Bellman. holds for all t ∈ I, then we obtain for all t ∈ I. At first, let us state the model's assumptions [16, 17] : 1) The population's size N is fixed over time, i.e., N(t) = N for all t ∈ [0, ∞); 2) We divide a population into three homogeneous subgroups, namely susceptible people (S), infectious people (I), and recovered people (R). We can clearly assign every individual to exactly one subgroup. Hence, we obtain for all t ∈ [0, ∞); 3) Additionally, no births and deaths occur; We briefly comment on our choice of time-varying transmission rates. The choice of timedependent transmission rates is possible because countermeasures such as lock-downs, social distancing, or other political actions such as curfews reduce possible contacts between susceptible and infectious people. In addition to that, time-varying recovery rates might also be an interesting choice due to different medical treatments over the course of new epidemics such as COVID-19. Furthermore, we exclude age structures or spatial relationships from our time-continuous model [16, 19] . For abbreviation, we write f (t) := df (t) dt . Our equations of the timecontinuous SIR model read as follows: with initial conditions S(0) = S 1 > 0, I(0) = I 1 > 0, and R(0) = R 1 ≥ 0. We portray a chart of the flow between the different three subgroups in Fig. 1 . Obviously, the equation is valid, and hence the first assumption is fulfilled. Now, we prove boundedness of the solution. For this purpose, we modify ideas from [17] and [44] because we, in contrast to them, consider bounded, time-varying transmission rates α : [0, ∞) − → [0, ∞) and recovery rates β : [0, ∞) − → [0, ∞). Proof Here, we split the proof into three parts. 1) We first consider S (t) = -α(t) · I(t)·S(t) N . Separation of variables leads to Integration yields and this implies Hence, it holds S(t) ≥ 0 for all t ≥ 0. 2) Let us continue with I (t) = α(t) · I(t)·S(t) N β(t) · I(t). Separation of variables gives us and, applying the same argument as in our first step, we conclude This yields I(t) ≥ 0 for all t ≥ 0. 3) Finally, since R (t) = β(t) · I(t) holds, we clearly obtain and R(t) ≥ 0 for all t ≥ 0 is valid. This completes our proof. Since N = S(t) + I(t) + R(t) holds by our first assumption, we can finally state our boundedness theorem. For all solution functions of (8), we have: In contrast to many other works, we formulate a theorem regarding global existence in time of (8) based on Theorem 1. For abbreviation, we introduce the supremum norm on an arbitrary time interval [a, b] for a continuous function f : [a, b] − → R. A similar definition holds for vector-valued functions. In our case, we consider [0, ∞) in general. Now, we show that a possible solution to (8) exists for all times t ≥ 0. The nonlinear first order ordinary differential equation system (8) has at least one solution which exists for all t ≥ 0. Proof By defining z(t) = (S(t), I(t), R(t)) T , we can set Clearly, G is Lipschitz continuous. By considering the supremum norm on our Euclidean space and applying the triangle inequality, we get from (9) by the boundedness of our solution functions and the boundedness of our timevarying transmission and recovery rates. Thus, all our assumptions of Theorem 1 are fulfilled and our proof is complete. We present our proof of global uniqueness in time based on an inductive application of Banach's fixed point theorem, i.e., that the initial value problem (8) is uniquely solvable for all times t ≥ 0. The nonlinear first order ordinary differential equation system (8) has a unique solution that exists for all t ≥ 0. Proof 1) Let us first consider the time interval [0, τ ] where we have to choose τ accordingly such that Banach's fixed point theorem is applicable. 2) We need one brief lemma. Let x 1 , x 2 , y 1 , y 2 ∈ R be arbitrary. By zero addition and application of the triangle inequality, we obtain by our inequality of the second step and application of the triangle inequality. 5) Furthermore, we conclude that holds. 6) Summarizing the previous steps, we obtain If we choose τ := 1 2·(2·α max +β max ) , this implies and hence we conclude the uniqueness of solution on the time interval [0, τ ]. 7) Inductively, we see that we can derive this contraction property on all time intervals [k · τ , (k + 1) · τ ] for all k ∈ N 0 by choosing k · τ as our starting point and for our initial conditions. Henceforth, our proof of global uniqueness in time is complete. Here, we consider the perturbed initial value problems with initial conditions S a (0) = S a,1 > 0, I a (0) = I a,1 > 0, R a (0) = R a,1 ≥ 0 and are different time-varying transmission and recovery rates. Now, we prove that small perturbations in initial conditions or small differences in time-varying transmission or recovery rates lead to small differences in the solutions on short time intervals [0, T]. This fact is summarized in the following theorem. ) T be the solutions of (10) and (11) . Define the function and the constant We see that holds for arbitrary t ∈ [0, T] with given T ≥ 0. Proof 1) Let us first mention that we often use the inequality for arbitrary x 1 , x 2 , y 1 , y 2 ∈ R as proven in Theorem 6. Additionally, we see that 2) At first, we obtain the inequality for arbitrary t ∈ [0, T] by application of the triangle inequality, boundedness of our all functions, and the inequality of our first step. 3) Secondly, we have to estimate |I a (t) -I b (t)|. We see that holds for arbitrary t ∈ [0, T]. The summand I can be estimated in the previous step. This For the third summand II, we observe that is valid for arbitrary t. Summarizing these results, we obtain Since all the assumptions of Theorem 3 are fulfilled, we see that holds for arbitrary t ∈ [0, T], which finishes our proof. We now investigate the long-time behavior of solution, some monotonicity properties and summarize our results in the following theorem. We get: 1) S is monotonically decreasing, and there exists a number S ≥ 0 such that lim t→∞ S(t) = S . It even holds S > 0; 2) R is monotonically increasing, and there exists a number R ≥ 0 such that lim t→∞ R(t) = R ; 3) I is Lebesgue-integrable on [0, ∞) and lim t→∞ I(t) = 0 for all solution functions of (8). is monotonically decreasing and bounded below by zero. This implies the existence of S ≥ 0 such that lim t→∞ S(t) = S . Additionally, by considering S (t) R (t) for t > 0, we obtain and separation of variables implies Integration yields is monotonically increasing and bounded above by N . This yields the are bounded and nonnegative. Therefore, we obtain that I is Lebesgue-integrable on [0, ∞) and lim t→∞ I(t) = 0. This finishes our proof. In our nonautonomous SIR model, the time-dependent basic reproduction number can be defined by which is similar to the constant case [17, 45, 46] . (13) is well defined. Proof We observe that is valid for all t ≥ 0. This proves our claim. In this section, we examine time-discrete versions of the given time-continuous SIR model (8) . Here, we only state a fully explicit scheme (14) and a fully implicit scheme of the time-continuous SIR model (8) for all j ∈ {1, . . . , M -1}. However, the fully explicit scheme (14) simply reduces to a linear system, while the fully implicit scheme (15) maintains the nonlinear structure of the continuous problem formulation (8) . For this reason, we investigate this fully implicit scheme in the following. We assume that 0 < α min ≤ a j ≤ α max < 1 and 0 < β min ≤ β j < β max ≤ 1 are given for all j ∈ {1, . . . , M} and that 0 < t j+1t j ≤ 1 for all j ∈ {1, . . . , M -1} and that S 1 > 0, I 1 > 0, and R 1 ≥ 0 are given. As later observed in our numerical examples, these assumptions are fulfilled in epidemiological data of the spread of COVID-19. An implicit solution scheme of (15) reads as follows: for all j ∈ {1, . . . , M -1}. Now, we are able to obtain an appropriate solution scheme from (17) which even implies unique solvability for all j ∈ {1, . . . , M -1} under the assumption that S j > 0, I j > 0, and R j ≥ 0 for all j ∈ {1, . . . , M}. Our main ingredient is the equation from (17) . Plugging (18) and writing j+1 = (t j+1t j ) yields for all j ∈ {1, . . . , M -1}. Hence, we get and by setting and we get A j+1 · I 2 j+1 + 2 · B j+1 · I j+1 = N · I j and can finally conclude for all j ∈ {1, . . . , M -1}. We now have an explicit solution formula for I j+1 for all j ∈ {1, . . . , M -1} and therefore also for S j+1 and R j+1 for all j ∈ {1, . . . , M -1}. Summarizing our results, we can formulate the following theorem. Theorem 9 Let us assume that 0 < α min ≤ a j ≤ α max < 1 and 0 < β min ≤ β j ≤ β max < 1 are given for all j ∈ {1, . . . , M}, that 0 < t j+1t j ≤ 1 holds for all j ∈ {1, . . . , M -1}, and that S 1 > 0, I 1 > 0, and R 1 ≥ 0 are prescribed. The implicit solution scheme (17) is uniquely solvable for all j ∈ {1, . . . , M -1}. It holds (22) from (20) and (21). We show that many of the continuous properties from Theorems 4 and 8 even translate to the time-discrete implicit scheme (17) . (17), we have: Proof 1) It holds I j ≥ 0 due to (22) and I j ≤ N due to (16) for all j ∈ {1, . . . , M}. 2) By our first property and due to (16) , we have the inequality 0 ≤ S j ≤ N for all j ∈ {1, . . . , M}. Again by our first property, we obtain 3) By our first property and due to (16) , we obtain the inequality 0 ≤ R j ≤ N for all j ∈ {1, . . . , M}. Again by our first property, we conclude 4) Since {R j } j∈N is monotonically increasing and bounded above by the total population size N , there exists a nonnegative constant R such that lim j→∞ R j = R . Furthermore, it holds R j+1 -R j = β j+1 · j+1 · I j+1 , which yields This implies lim j→∞ I j = 0 and completes our assertion's proof. Now, we want to provide an upper bound for error propagation. Before proving this statement, we need to formulate some assumptions for our convergence analysis. We summarize them in the following list: nonnegative constants α min , α max , β min , β max such that 0 < α min ≤ α(t) ≤ α max < 1 and 0 < β min ≤ β(t) ≤ β max < 1 hold for all t ∈ [0, T]; 6) Choose p < min{ 1 4·(α max +β max ) , 1} ≤ 1 for all p ∈ N and set := max p∈N p . Under these conditions, we obtain the following theorem where we adapt ideas from the error analysis of an explicit-implicit solution algorithm as presented in [20] . Proof We briefly describe our strategy first because this proof is technical. We begin with an estimation of local errors between time-continuous and time-discrete solutions. Afterwards, we consider error propagation in time. Conclusively, we investigate the cumulation of these errors. Time-discrete solutions are written as S p at time t p and time-continuous solutions as S(t p ) at the same time. 1) For examination of local errors, we assume that Here, we consider solely one time step and denote corresponding time-discrete solutions by S p+1 , I p+1 , and R p+1 . , and this implies :=II S,1 by application of the triangle inequality. We estimate these terms separately. For I S,1 , we obtain and the mean value theorem of calculus implies the existence of ξ S,1 ∈ (t p , t p+1 ) such that holds. This yields For II S,1 , we see that :=III S,1 is valid by the definition of S (t), boundedness of our solution functions, and application of the triangle inequality. By plugging into III S,1 , we obtain by the boundedness of our solution functions and application of the triangle inequality. By the mean value theorem, there exists ξ α,1 ∈ [t p , t p+1 ] such that holds. Hence, an additional application of the triangle inequality and boundedness of the solution functions yields Plugging (25) into II S,1 , we conclude that holds. Combining (24) and (26), we infer that is valid. Hence, it follows by application of the triangle inequality and with similar arguments as provided in the previous step. As holds, we further obtain by the boundedness of the solution functions and application of the triangle inequality. By we obtain by the boundedness of the solution functions, the mean value theorem of calculus, and application of the triangle inequality. Additionally, it holds by application of the mean value theorem of calculus. Combining (28) and (29) and plugging these results into Plugging this inequality into holds. This implies By applying R (t p ) = β p · I(t p ) and Plugging this inequality into 1.4) Define C loc := max{C loc,S ; C loc,I ; C loc,R }. It holds for local errors on time intervals [t p , t p+1 ]. 2) In reality, (t p , S p ) T , (t p , I p ) T , and (t p , R p ) T do not exactly lie on the graph of the timecontinuous solution. Therefore, we must examine how procedural errors such as S p -S(t p ), I p -I(t p ) or R p -R(t p ) propagate to the (p + 1)th time step. These investigations are carried out in step 2) and in step 3). By (9) , we see that G(t p+1 , z p+1 ) -G t p+1 , z(t p+1 ) holds, and this implies 1 1 -2 · (α max + β max ) · j = C loc · 2 · ( 1 1-2·(α max +β max )· ) p -1 ( which finishes our proof of (23). In our nonautonomous time-discrete SIR model, the time-dependent basic reproduction number can be defined by for arbitrary k ∈ {1, . . . , M}, which is similar to the case of constant transmission and recovery rates [17, 45] . Proof This proof is identical to Lemma 2. We are now able to give a brief description of our numerical algorithm to solve the timediscrete implicit solution scheme (17) . Here, we summarize our inputs, our computational steps, and our algorithmic outputs. We sketch the resulting algorithm in Table 1 . We apply our time-discrete implicit SIR solution scheme (22) from Table 1 to available data regarding the spread of COVID-19 in Germany and Iran from John Hopkins University [1, 2] . These countries are chosen because they update confirmed, dead, and estimated recovered cases on a regular basis. In Table 2 , we summarize projected population sizes for 2019 from the United Nations [47] . Step 1: Step 2: -ComputeI j+1 by (22) , (21) and (20) At first, we consider the example of Germany in detail. We thoroughly describe our approach to check our model's validity. We only give some simple parameter estimation techniques and vary user-chosen time-dependent parameter functions. Since inverse problems are an active field of research, we refer the readers to works by Bock and Schittkowski for more sophisticated parameter estimation techniques in dynamical systems [48, 49] . Our work also focuses on usefulness in possibly describing real-world data. Afterwards, we state some computational results for data from Iran. for all j ∈ {1, . . . , M}. The unprocessed and processed data for Germany are depicted in Figs. 2 and 3. On the one hand, it can be clearly seen in Fig. 2 that both sequences of cumulative infected and cumulative recovered people are monotonically increasing for our unprocessed data. On the other hand, we notice in Fig. 3 that the behavior changes for our processed data. Here, we present an algorithm to calculate our time-varying transmission and recovery rates based on our numerical algorithm (15) for all j ∈ {1, . . . , M -1}. We rely on the first Summarizing our results, we obtain and and a short algorithmic summary can be found in Table 3 . The time-varying transmission rate from real-world data for Germany is presented in Fig. 4 . Clearly, the transmission rate decreases due to countermeasures such as local lock- Step 1: Step 2: -F orallj ∈ {2, . . . , M}, compute α j and β j according to (37) and (38) The time-varying recovery rate from real-world data for Germany is depicted in Fig. 5 . At the early stage of an epidemic, there are possible just few recoveries, thus this rate is relatively small. After some time, this situation changes as more people defeat the disease and recover. The rate seems to be constant with heavy variations due to the test capacity. Additionally, there are unknown cases because these people might have a mild disease course. Now, the time-dependent basic reproduction number R 0 (t j ) is readily computed by (35) . Our computational results from this approach are portrayed in Fig. 6 . Since there are only few recovered people at the beginning of disease, our computations provide high numerical basic reproduction number. We observe that the computational basic reproduction number is monotonically decreasing in spring due to political countermeasures and so- cial distancing. However, the graph shows that the computation basic reproduction number rises in summer because contacts between people rose. Variations are seen for similar reasons as mentioned for our time-dependent transmission and recovery rates. We only briefly sketch the parameter identification problem because it is an inverse problem [50, 51] . A deep discussion is beyond the scope of this paper and would be a topic of Step 1: -Compute β by (43) Step 2: -Compute γ 2 by (45) Step 3: -Compute γ 1 by (44) Step 4: -Compute α 1 and α 2 according to transformation (41) Outputs: -Parameters α 1 , α 2 , and β for our parametric rates (39) and (40) By the Hölder inequality, we obtain with equality only in the case that all t j are equal. Since we have a strictly increasing time sequence, the second determinant of the upper-left sub-matrices is also positive. Finally, the determinant of the full matrix is positive as well. Hence, our cost function J is strictly convex. Conclusively, it possesses a unique local minimizer by [52, Theorem 2.4]. 2) By strict convexity, we infer that the unique local minimizer is also the unique global minimizer of our cost function J by [52, Theorem 2.5]. We summarize our algorithmic approach for parameter estimation of our time-varying transmission and recovery rates in Table 4 . predictions In Fig. 7 , we see that the assumption of exponentially decaying time-dependent transmission rates is acceptable at the beginning of spreading disease with respect to German data. Due to short-term prediction, we notice that the constant recovery rate is underestimated in Fig. 8 . Conclusively, both assumptions seem to be acceptable at the first weeks of a spreading disease. Computational results for two models on the time interval [25, 62] are depicted in Figs. 9-12. Figures 9-12 indicate that sensitivity of parameters is really an issue in epidemiological models. This is in accordance with Theorem 7. These results also imply that an exponentially decaying transmission rate is an acceptable choice at the beginning of spreading disease. Figures 4 and 5 indicate that constant transmission and recovery rates are reasonable assumption at later stages. Computational results can be found in Figs. 13 and 14 . We also notice that the number of infected people rises in summer, possibly due to more social contacts. This could eventually be regarded as the beginning of a 'second wave' . For future investigations, it might be interesting to use more complex transmission rates or piecewise defined functions with switching points, see e.g. [49] . Recovery rates from real-world data and from parameter estimation for short-term prediction of German data with t 1 = 0 (1 March 2020) and t M = 62 (2 May 2020). The first estimated recovery rate reads β ≈ 0.04403 for the mean value on the full interval. The second estimated recovery rate reads β ≈ 0.063 as the mean value on the time interval [25, 62] because the fluctuations in β arise while there is no recovery rate estimation possible for the first days Real-world data from Iran and short-term computational results for COVID-19 data from Iran can be found in Figs. 15-23. Figure 17 again supports the assumption of an exponen- These results indicate that time-dependent transmission rates are a necessary addition to the classical SIR model. Alternatively, models with fractional derivatives could be considered [53, 54] . We established certain properties such as well-posedness of the solution of our timecontinuous SIR model in Sect. 2. Fortunately, we were able to transfer many properties of the time-continuous model to our time-discrete implicit SIR model in Sect. 3. These include unique solvability and monotonicity properties. In contrast to many other works Computational results for our model of recovered people in Germany for user-chosen α 1 ≈ 0.040, α 2 ≈ 0.0, β ≈ 0.075 for short-time simulation on t 1 = 61 (1 May 2020) and t M = 184 (1 September 2020) with real-world data as initial conditions mentioned in Sect. 1, we avoid an explicit forward model, but we could transform our implicit scheme to an easily solvable scheme. Thus, this makes our proposed scheme an attractive first prediction choice. In addition to that, we showed that our numerical scheme possesses an upper error bound. Regarding our computational results, we see that our parametrization is an appropriate fit for first forecasts considering the first wave of a spreading virus. Since these transmission rates are monotonically decreasing, we, however, remark that we will need to use another parametrization if we want to model diseases with seasonal behavior [55] . Regarding our chosen examples, we get reasonable results. Additionally, we observe that our theoretical findings regarding monotonicity of recovered people from The- As depicted in Sect. 4, the inverse problem definitely needs further investigation. This is a topic of its own interest [56] [57] [58] [59] since we need tools from different mathematical disciplines. As future research directions, extensions to further epidemiological forward models should be considered as we surely need more tools to predict the impact of upcoming epidemics. One can also consider delayed-differential or stochastic variants of our SIR model or modifications and extensions [23, 27] because, from a biological point of view, we often have to face integration of incubation times in epidemic models. An interactive web-based dashboard to track COVID-19 in real time Hopkins University (dataset): novel coronavirus (COVID-19) cases data A pneumonia outbreak associated with a new coronavirus of probable bat origin Clinical characteristics and intrauterine vertical transmission potential of COVID-19 infection in nine pregnant women: a retrospective review of medical records Pathological findings of COVID-19 associated with acute respiratory distress syndrome Risk of COVID-19 for patients with cancer Aerodynamic analysis of SARS-CoV-2 in two Wuhan hospitals COVID-19-associated acute hemorrhagic necrotizing encephalopathy: CT and MRI features Acute pulmonary embolism associated with COVID-19 pneumonia detected by pulmonary CT angiography A contribution to the mathematical theory of epidemics Contributions to the mathematical theory of epidemics-I Contributions to the mathematical theory of epidemics-II. The problem of endemicity Contributions to the mathematical theory of epidemics-III. Further studies of the problem of endemicity The mathematics of infectious diseases Mathematical Models in Population Biology and Epidemiology An Introduction to Mathematical Epidemiology The SIS-model on time scales The Basic Approach to Age-Structured Population Dynamics: Models, Methods and Numerics An age-and sex-structured SIR model: theory and an explicit-implicit numerical solution algorithm Inferring COVID-19 spreading rates and potential change points for number forecasts Analysis and forecast of COVID-19 spreading in China, Italy and France A discrete stochastic model of the COVID-19 outbreak: forecast and control Report 13: estimating the number of infections and the impact of non-pharmaceutical interventions on COVID-19 in 11 European countries Estimating the number of infections and the impact of non-pharmaceutical interventions on COVID-19 in European countries: technical description update Effective containment explains sub-exponential growth in confirmed cases of recent COVID-19 outbreak in Mainland China A stochastic differential equation SIS epidemic model Forecasting seasonal influenza with a state-space SIR model Optimization method for forecasting confirmed cases of COVID-19 in China The role of power decay, exponential decay and Mittag-Leffler function's waiting time distribution: application of cancer spread Modelling the spread of COVID-19 with new fractional-fractal operators: can the lockdown save mankind before vaccination? Existence of solution and stability for the fractional order novel coronavirus (nCoV-2019) model Modeling the dynamics of novel coronavirus (2019-nCov) with fractional derivative A new fractional-order compartmental disease model The Kermack-McKendrick epidemical model revisited Some simple epidemic models Exact solution to a dynamic SIR model Some time-discrete SI, SIR and SIS epidemic models Ordinary Differential Equations: Basics and Beyond Revisiting maximum log-likelihood parameter estimation for two-parameter Weibull distributions Nonlinear Function Analysis and Its Applications I-Fixed-Point Theorem Functional Analysis Inequalities for Differential and Integral Equations Exact analytics solutions of the susceptible-infected-recovered (SIR) epidemic model and of the SIR model with equal death and birth rates The concept of R 0 in epidemic history Reproduction numbers of infectious disease models Population Dynamics: World population prospects 2019 (total population-both sexes Model Based Parameter Estimation-Theory and Applications Springer Numerical Data Fitting in Dynamical Systems Inverse Problem Theory and Methods for Parameter Estimation Inverse problem for coefficient identification in SIR epidemic models Numerical Optimization The Analysis of Fractional Differential Equation Numerical Methods for Fractional Differentiation Estimating influenza deaths in Canada The inverse problem in mathematical biology The parameter identification problem for SIR epidemic models: identifying unreported cases A time delay dynamical model for outbreak of 2019-nCoV and the parameter identification Time-discrete parameter identification algorithms for two deterministic epidemiological models applied to the spread of COVID-19 Both authors acknowledge support of the Max-Planck-Institute for Dynamics and Self-Organization such that they could finish their research on this work. In addition to that, the authors thank both referees for their valuable comments which definitely helped us to improve our manuscript's quality. Both authors thank the Max-Planck-Society for covering the article processing charge. CSV-files with data for confirmed cases, death cases, and recovered cases from different countries, including Germany and Iran, are available in [2] . The files are named as follows: 1) TIME_SERIES_COVID19_CONFIRMED_GLOBAL.CSV; 2) TIME_SERIES_COVID19_DEATHS_GLOBAL.CSV; 3) TIME_SERIES_COVID19_RECOVERED_GLOBAL.CSV, and many data from different countries can be found within these CSV-files. Population data for different countries can be checked in [47] . More precisely, readers should check the file for total population with both sexes by the standard projection methods of the United Nations Department of Economic and Social Affairs. The authors declare that they have no competing interests. BW and JS designed the research. BW analyzed time-continuous and time-discrete models. BW implemented the time-discrete version. BW and JS analyzed data, discussed results, and presented the data. BW and JS drafted and edited the manuscript. Both authors read and approved the final manuscript. Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. We see thatholds, and this impliesHence, we concludewith := max p∈{1,...,M-1} p+1 < 1 4·(α max +β max ) . 3) Now, we want to prove the upper error bound between the time-discrete solution and the time-continuous solution. At first, we notice thatis valid for p = 1 by (32), by (33) , and by our assumption that initial conditions of the timecontinuous and the time-discrete models coincide. For p = 2, we obtainFor arbitrary p ∈ {1, . . . , M -2}, we assume thatby induction. By applying the geometric series, we obtainown interest. By looking at Figs. 4 and 5, we assumeandwith real constants α 1 , α 2 , and β which we determine from real-world transmission and recovery rate sequences { α j } M j=2 and { β j } M j=2 . Since α j > 0 and β j > 0 for all j ∈ {2, . . . , M}, we can assume α 1 > 0 and β > 0. Since α(t) is nonlinear, we use the transformationwith γ 1 := ln(α 1 ) and γ 2 := -α 2 as in the case of maximum log-likelihood estimation. Now, our cost function J : R 3 − → [0, ∞) reads as follows:We obtain the following theorem. Proof 1) We first show that J possesses a unique local minimizer.1.1) To achieve our goal, we calculate the first derivatives. We obtain{ββ j }. To get a local minimizer ( γ 1 , γ 2 , β), it is necessary that all partial derivatives vanish at candidates for local extrema. 1.2b) Setting ∂J ∂γ 1 ( γ 1 , γ 2 , β) = 0, we infer thatholds. 1.2c) If we set ∂J ∂γ 2 ( γ 1 , γ 2 , β) = 0, we obtaint j and this yieldsFinally, we conclude thatholds.