Understanding Hematopoietic Stem Cell Dynamics—Insights from Mathematical Modelling

In this section, we supplement recent reviews of mathematical modelling of stem cell function written by Stiehl and Marciniak-Czochra [16], and Brunetti, Mackey, and Craig [17].

In 1978, Mackey proposed a mathematical model [18] which inspired many of the mathematical models of blood cell formation that we describe in this article. The model proposed by Mackey describes the dynamics of a population of pluripotent stem cells, explicitly considering the cell cycle with HSCs shifting between a proliferating phase and a \(_\) resting phase. Analyzing the model, the author connects changes in stem cell proliferation with the production of blood cells. The author focuses on a group of specific disorders characterized by periodic oscillations of mature blood cell counts. Mathematical models of this class of diseases are also reviewed by Dale and Mackey in 2015 [19]. The early paper by Mackey is an illustrative example of how mathematical modelling can be used to relate cell properties, e.g., the rate of apoptosis to dynamics of the entire pool of HSCs.

In a later paper [20], Mackey shows that the equivalent of the total human body weight worth of blood cells is produced about every 7 years in the typical human adult. In the same paper, the author relates an extended version of his previous model to data from an HSC-labelling experiment, estimating the differentiation rate of HSCs in humans to be between 0.01 and 0.02 days−1 and the apoptosis rate between 0.07 and 0.23 days−1 in a steady-state situation, i.e., in a situation where cell counts remain constant. In a similar effort to understand HSC processes, Abkowitz and colleagues [2124] relate stochastic models of HSCs to in vivo data of murine, feline, and human origin, and are able to obtain estimates of, e.g., the replication rate of human HSCs. Furthermore, the authors suggest that the number of HSCs is comparable across different mammals, with estimates for humans observed in the blood, the success is onlyin the range of 11,400 to 22,400 [23]. Based on this work and [22], Catlin et al. [24] use 11,000 HSCs as an estimate for the steady-state size of the HSC pool, a number which has been used subsequently in several other models of HSC dynamics. Further examples of the application of mathematical models to infer cell kinetics based on labelling experiments in animals can be found in [25••] and [26].

Another early approach to model HSC dynamics is proposed in the work of Dingli and Michor [27]. In this model, hematopoiesis is described by a system of ordinary differential equations, in which one of the variables represents the pool of HSCs. The authors consider a rate of self-renewing production of HSCs, a rate of apoptosis as well as a rate of production of differentiated mature blood cells. The model is then extended for a population of leukemic stem cells (LSCs) with similar behavior, where division rates of the healthy HSCs and the LSCs are negatively regulated by the total number of stem cells. The resulting model captures how healthy hematopoiesis is suppressed by an increase in LSCs as hematologic malignancy progresses. The most significant result of the mathematical analysis is provided in the title: “Successful therapy must eradicate cancer stem cells” [27]. The high genetic similarity of diagnosis and relapse samples in acute myeloid leukemia, a severe blood cancer, supports this claim. Observations in sequencing studies are in line with the concept that the relapse is either triggered by cancer stem cells which have survived therapy or by malignant or pre-malignant stem cells which have acquired additional mutations [54]. This suggests that while some leukemia therapies may be successful at reducing the disease burden observed in the blood, the success is only temporary if they do not act on the population of HSCs and LSCs. This conclusion has later been confirmed by related models [6, 28]. In their later work, Dingli and coworkers propose models of hematopoiesis which account for multiple immature cell states [29, 30]. These models are used to predict the number of mitotic events linking the stem cell state to the mature cell compartment and to infer the phenotypic consequences of a mutation linked to oscillating blood cell counts.

Another approach to the mathematical modelling of HSCs is described in the model proposed by Roeder and Loeffler [31] and further investigated together with coworkers [3234]. In the model, cells are assumed to move between two distinct growth environments, named GE-A and GE-Ω. In one of these abstract environments, GE-Ω, the cells are actively cycling and dividing while they are quiescent (non-dividing) in the other environment, GE-A. Each cell is characterized by its cycling status and its affinity to the GE-A environment. In contrast to other HSC models where a predefined pool of HSCs divides according to a model recapitulating the classical cell cycle phases, the model of Roeder and Loeffler [31] considers stemness and cycling activity as a property that emerges from cells dynamically switching between the two environments. The switching of a given cell between the environments is described as a function of its affinity to the GE-A environment and of the cell counts. Affinity is preserved or raised during quiescence while division degrades affinity. When affinity is sufficiently small, the cell is considered as terminally differentiated and unable to return to quiescence. Roeder and Loeffler [31] show through model simulations that the model is consistent with various experimental results from the literature, including those discussed by Abkowitz et al. [21]. Hence, the model provides a framework to understand experimental data, with the perspective that one should not only consider cell stemness as intrinsic to the cells but rather consider the interaction between HSCs and their environment. The notion of HSC reversibly exiting and re-entering quiescence, possibly adapting dynamically to external signalling, is in agreement with the notion of bone marrow stem cell niches, i.e., specific micro-environments supporting HSC quiescence and repair. A thorough review of the biological research on HSC niches is given by Wilson and Trumpp [35]. Trumpp, Essers, and Wilson [36] suggest that hematologic malignancies could be efficiently treated by combination therapy in which one drug stimulates the activation of quiescent HSCs while another drug then eradicates the newly activated non-quiescent HSCs. This is later investigated in the work of Glauche et al. [37] by adapting the model from [31], providing a mechanistic explanation behind the efficacy of combination therapy of chronic myeloid leukemia (CML).

A recent study by Ashcroft and colleagues models the binding of HSCs to HSC-specific niches in the bone marrow [38••]. In the model, HSCs detach from the niche and enter the peripheral blood at a certain rate. Upon returning to the bone marrow, the HSCs can reattach to unoccupied niches. Natural death of HSCs is assumed to occur more frequently in the peripheral blood, while only niche-bound HSCs divide. In the model, the majority of HSCs are assumed to be attached to a niche and quiescent, with a low number of unoccupied niches available in steady-state hematopoiesis. While Ashcroft et al. specifically relate the model to murine data, the model and related results should be valid for human HSCs as well. By using ordinary and stochastic differential equations, the authors investigate the prerequisites for clonal dominance in the stem cell niche. They conclude that clonal dominance in mice requires a selective advantage and cannot be the result of neutral drift. Furthermore, their model is used to investigate HSC dynamics after bone marrow transplantation. In this context, transplantation with and without prior emptying of the niches by chemotherapy (so-called preconditioning) is considered. Following the transplantation of HSCs without selective advantage into the blood stream, the engraftment of the transplanted cells in the bone marrow is limited by the number of unoccupied niches. The model can explain the experimental observation from Bhattacharya et al. [39, 40] that multiple smaller transplantations over a few days can lead to higher uptake of HSCs into the murine bone marrow compartment compared to a single-bolus transplant [38••]. The authors also use the model to predict the chimerism (abundance of donor-derived cells) after the transplantation of cells without selective advantage. Furthermore, Ashcroft et al. [38••] investigate how the probability of reconstitution after preconditioning depends on the transplanted stem cell dose. Special attention is paid to the scenario where only a single cell is transplanted. Experimental verification of such scenarios can be challenging, although not entirely impossible [41]. This is an example of an experiment that can easily be carried out in silico in a mathematical model. A related example is how patients, which in reality can only receive one treatment at a time, would have responded to a different treatment protocol or a different stem cell dose [42]. Other models of bone marrow transplantation include [43, 44] for the human and [45••] for the murine case.

Based on the observation that HSC counts are reduced in many acute myeloid leukemia (AML) patients [46], Stiehl et al. and Wang et al. [9••, 46] propose a model where HSCs and LSCs share a joint stem cell niche. This model is an extension of previous models which have been shown to recapitulate clinical blood sample data over time following bone marrow transplantation [42, 43] and the growth of leukemic cells [8, 9••, 10]. The model is given by a system of ordinary differential equations. Stem cell dynamics in the niche are coupled to the time evolution of progenitor, precursor, and mature cells as well as leukemic blasts. Clonal heterogeneity of HSCs and in turn the expansion of, e.g., a leukemic clone in the model of [38••] requires the presence of unoccupied niches and spontaneous detachment of HSCs from the niches. The model proposed in [9••, 46] considers a scenario where LSCs can actively dislodge HSCs from the niche. Since stem cells require the niche to maintain stemness properties, an invading clone has to conquer niche spaces to expand. Upon division of a LSC one of the two progeny occupies the niche of the parent, whereas the other attempts to dislodge an HSC from the niche. If this is successful, the dislodged HSC differentiates and the LSC maintains its stemness. In the opposite case, i.e., after a certain number of futile dislodgement attempts, the LSC differentiates. Similar assumptions hold for HSCs; however, to observe the expansion of leukemic cells, the probability that LSCs dislodge HSCs must outweigh the probability that HSCs dislodge LSCs. A leukemic clone with a sufficiently pronounced ability to dislodge HSCs is able to take over the niche, while the blood production might initially only show minor signs of the disease since it is transiently maintained by long-term progenitors. This model can explain the observation from [46] that AML patients with low HSC frequencies (in [46] defined as HSC frequencies below 15% of the median HSC frequency of healthy human individuals) have a poor survival. Furthermore, it is in line with the observation that HSC counts decrease before overt relapse of the AML [46]. Models assuming that HSCs and LSCs reside in separate niches and interact via systemic feedbacks that depend on the mature cell and leukemic blast counts cannot reproduce these observations [46]. A quantitative version of this model can be used to understand how the competition of stem cells for niche spaces impacts the clinical course of AML [9••, 11]. The authors develop a model-based risk stratification which uses HSC and blast counts at the time of AML diagnosis [9••] and show that this approach allows to refine the prognostic scoring of the patient cohort from [46].

In most of the models discussed above, the number of HSC niches is assumed to be independent of the dynamics of the HSC pool. To investigate the effect of feedback from HSCs to the niche-forming cells, [47] explicitly model the population of niche cells. Under the assumptions that niche cells undergo apoptosis in the absence of HSCs and that contact of HSCs to niche cells induces quiescence of HSCs but proliferation of niche cells, the model exhibits a homeostatic state. After perturbations, the model returns to homeostasis in a dampened oscillatory manner reminiscent of integral feedback control. The work of [47] illustrates that explicit modelling of the cells constituting the BM-niche can provide additional insights into the regulations of HSC dynamics.

Finally, recent work by the authors of this review aimed to model competition between HSC clones through niche occupation. Inspired by many of the models described above, a model was proposed in which competition for a limited pool of niches naturally leads to an expression of HSC fitness [48]. When two clones compete for the same niche space, e.g., in a situation where a leukemic clone has arisen due to mutations, the clone-specific fitness determines which of the two clones will eventually out-compete the other. In a scenario where the two clones have the exact same fitness, the clones will co-exist indefinitely. The model provides an expression of clonal fitness in terms of stem cell properties such as rates of proliferation, differentiation, and niche attachment. Combining this HSC model with a previous model of myeloproliferative neoplasms (MPNs) developed by Andersen et al. [49] suggests that differences in HSC fitness can explain both the disease-progression of MPNs as well as the efficacy of certain treatment protocols [50].

As approximations of reality models naturally have limitations. Besides limitations originating from the mathematical assumptions of the used modelling frameworks such as the high computational costs in the case of individual-based models or the neglect of random events in ordinary differential equation, there exist limitations coming from gaps in biological knowledge or from simplifying assumptions made during model design. One common source of limitations is the sparsity of in vivo data, especially in the human case. This sparsity originates from the inaccessibility of relevant quantities, such as kinetic parameters of stem cells, or the small number of follow-up time points. The sparsity of the data can lead to the unidentifiability of model parameters and, therefore, make it impossible to quantify important processes. Due to the sparsity of the data, it can also be challenging to distinguish between competing models. Since most human cancers require an instantaneous start of the therapy, longitudinal data of untreated cancers is practically unavailable. This makes it challenging to validate or to parameterize models of cancer dynamics. In most cases, invasive sampling techniques such as biopsies are required to extract the relevant information about a cancer. Since usually healthy individuals are not subjected to these invasive procedures, data on cancer evolution before clinical disease manifestation is rarely available. Therefore, a large part of our knowledge on tumor evolution is derived from experimental data or has been obtained by applying modelling or heuristic techniques to samples obtained at the time of diagnosis. Furthermore, important parts of our pathophysiological and pharmacological knowledge have been gained using animal or in vitro models, which do not necessarily reflect the human system [51, 52]. Although the latter may improve in the future due to advanced organoid or microfluidic systems, it has to be taken into account that human diseases are more complex than specific experimental systems [51, 52]. These limitations make it vital to validate models using real-world data and to be aware of the simplifying assumptions that have been applied during model development. In the context of personalized medicine, the applicability of models in clinical routine is further limited by the large inter-individual heterogeneity of the disease evolution, which is impacted by potentially unknown disease-specific factors (e.g., key mutations), patient-specific factors and environmental factors. Partially, this heterogeneity can be incorporated by the choice of individual model parameters for each patient [8, 9••, 11]. This, however, requires that the individualized parameters can be estimated based on patient data or derived from statistical or other approaches. Integration of mechanistic and artificial intelligence models is a promising direction to improve personalized medicine [53].

留言 (0)

沒有登入
gif