• No results found

University of Groningen On the molecular mechanisms of hematopoietic stem cell aging Lazare, Seka Simone

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen On the molecular mechanisms of hematopoietic stem cell aging Lazare, Seka Simone"

Copied!
139
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

University of Groningen

On the molecular mechanisms of hematopoietic stem cell aging Lazare, Seka Simone

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date: 2018

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Lazare, S. S. (2018). On the molecular mechanisms of hematopoietic stem cell aging. University of Groningen.

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

On the molecular mechanisms of

Hematopoietic Stem Cell aging

(3)

Cover design by Seka Lazare

ISBN (print version): 978-94-034-0419-6 ISBN (digital version): 978-94-034-0420-2 Copyright © 2018 by Seka Lazare

No parts of this book may be reproduced or transmitted in any form or by any means without permission of the author. All rights reserved.

(4)

On the molecular mechanisms of

Hematopoietic Stem Cell aging

PhD thesis

to obtain the degree of PhD at the University of Groningen

on the authority of the Rector Magnificus Prof. E. Sterken

and in accordance with the decision by the College of Deans. This thesis will be defended in public on Monday 12 February 2018 at 12.45 hours

by

Seka Simone Lazare

born on 4 June 1988 in Roseau, Dominica

(5)

Prof. G. de Haan

Co-supervisor

Dr. L.V. Bystrykh

Assessment Committee

Prof. G.A. Huls Prof. J.N.J. Philipsen Prof. D. Bryder

(6)

To my parents, Alick and Myrtle Lazare and my partner, James Corne

(7)

Jakub Wudarski Erik Zwart

(8)

Contents

CHAPTER 1 3

General Introduction

Aging of Hematopoietic Stem cells Blood 2017, in press

CHAPTER 2 23

Genomic and functional integrity of the hematopoietic system requires tolerance of oxidative DNA lesions

Exp Hematol. 2017 53:26-30

CHAPTER 3 59

Lifelong dietary intervention does not affect hematopoietic stem cell function Exp Hematol. 2017 53:26-30

CHAPTER 4 67

A comprehensive analysis of differentially expressed transcripts during hematopoietic stem cell aging in the mouse

Manuscript in preparation

CHAPTER 5 91

The Neogenin Receptor in Hematopoiesis Manuscript in preparation

CHAPTER 6 109

Summary and Discussion

APPENDICES 117

Dutch Summary/Nederlandse Samenvatting 118

Acknowledgements 124

Curriculum Vitae 129

(9)
(10)

CHAPTER 1

General Introduction

Aging of Hematopoietic Stem cells

Gerald de Haan* and Seka Lazare

European Research Institute for the Biology of Ageing, University Medical Center Groningen, University of Groningen. Antonius Deusinglaan 1, 9713 AV Groningen, the Netherlands.

*Corresponding author: g.de.haan@umcg.nl Blood 2017, in press

(11)

Abstract

Hematopoietic stem cells (HSCs) ensure a balanced production of all blood cells throughout life. As they age, HSCs gradually lose their self-renewal and regenerative potential, while the occurrence of cellular derailment strongly increases. Here we review our current understanding of the molecular mechanisms that contribute to HSC aging. We argue that most of the causes that underlie HSC aging result from cell-intrinsic pathways, and reflect on which aspects of the aging process may be reversible. As many hematological pathologies are strongly age-associated, strategies to intervene in aspects of the stem cell aging process may have significant clinical relevance.

Introduction

The average human life expectancy has increased very consistently during the last 150 years. The absolute number of elderly people has increased very substantially in many societies, and this is not likely to come to an end any time soon. Although it is evident that many elderly people reach advanced ages in healthy conditions, the prevalence of age-dependent disease and the average age of patients who enter the clinic is increasing. This is also true for patients that suffer from hematological diseases, as many hematological conditions are strongly age-dependent. From a clinical perspective, this raises two related, yet distinct, considerations. First, how does the hematopoietic system change with age, how does this lead to a spectrum of hematological conditions, and would it be possible to pharmacologically intervene in this process? Second, if hematological disease presents in an elderly patient, should this be treated differently than if it had occurred in a young patient? Although the latter is also of very significant clinical interest1 this manuscript will focus on our

understanding of the molecular mechanisms that make hematopoietic stem cells age. Only if we understand how hematopoietic stem cells age, we can begin exploring opportunities to prevent, delay, or even reverse aspects of the aging process.

HSC self-renewal and aging

Whereas research on aging has for the longest time been relatively descriptive, much progress has been made in the last decade to uncover the molecular drivers of biological aging. An influential review paper which describes the ‘Hallmarks of Aging’ provides a comprehensive overview of the various pathways that are believed to result in age-dependent cellular and organismal functional decline 2. One of these so-called

hallmarks of aging relates to stem cell exhaustion. Stem cell exhaustion refers to the gradual functional decline of adult, tissue-specific stem cells to maintain homeostasis of the tissue in which they reside. Stem cells are critically defined by their ability to self-renew, a concept which seems difficult to reconcile with the notion of aging; either stem cells are able to self-renew and therefore do not age, or alternatively, stem cells age, and therefore do not truly self-renew.

In the hematopoietic system, there are in fact multiple indications that stem cells do not formally self-renew, or at least have a restricted self-renewal potential and

(12)

are therefore fundamentally different from pluripotent ES or iPS cells. What are these indications? Historically, experimental hematologists have carried out serial stem cell transplantations in mice to assess the potential of these cells to multiply in vivo in myeloablatively conditioned recipients 3. Although serial transplantations

are obviously quite artificial (yet, they have also been performed in rare patient 4)

collectively these studies document that after transplant the hematopoietic stem

cell compartment never returns to normal values 5-7. If stem cells are submitted

to increased proliferative stress (by transplanting a low number, or even a single stem cell 8 or by initiating serial transplantations in short intervals) 9, self-renewal

is impeded more severely. Most notably, if stem cells that were first harvested from an aged donor mouse are serially transplanted in young recipients, the self-renewal capacity is much less compared to stem cells isolated from young donor mice 10. If

stem cells from a short-lived mouse strain are competed with stem cells from a long-lived mouse strain, the latter stem cells outcompete their shorter-long-lived counterparts

11. In general, in all studies in which young hematopoietic stem cells were competed

in transplantation studies against aged stem cells, without exception the young stem cells are functionally superior 12-16. These young stem cells produce, at the single cell

level more mature peripheral blood cells 10, 14 and they are better able to produce in a

balanced manner both myeloid and lymphoid cells 10, 17.

There is ample evidence that during steady state conditions the most primitive hematopoietic stem cells are very quiescent, and only divide very rarely. The concept of hematopoietic stem cell quiescence emerges from studies that have shown that stem cells are refractory to cell-cycle specific cytotoxic drugs, such as 5-fluorouracil

18 that stem cells are not easily labeled with BrdU 19, 20 and that it can take many

months before they contribute to hematopoiesis after transplant 21, 22. Very recently,

an elegant study showed that the most primitive of all hematopoietic stem cells may

undergo only 4 to 5 divisions in the lifetime of a mouse 23. This study suggested

that hematopoietic stem cells possess some sort of cellular memory, and the typical age-dependent phenotype of hematopoietic stem cells only emerges after they have divided 5 times. A very similar concept had been hypothesized to exist by one

of us more than two decades ago 24. Collectively, this indicates that most of the

proliferative burden resides with committed progenitors, and that during steady state hematopoiesis the most primitive stem cells are largely inactive. It seems plausible that with each cell division the potential of a hematopoietic stem cell to contribute to blood cell production declines, and that simultaneously the pool of stem cells with reduced potential increases, to compensate for loss of function of individual stem cells (Figure 1, adapted from 24, 25).

Most of what we know about HSC aging is based on studies that have used the mouse as an experimental model. As a consequence, it is not fully clear to what extent these molecular mechanisms also play a role in human hematopoietic stem cell aging. However, decades of research have shown that conceptually, the formation of blood cells in mouse and human is identical, and it is very plausible that mechanisms

(13)

that contribute to stem cell aging in mice, also do so in human. Indeed, it has been demonstrated that telomeres progressively shorten in human HSCs isolated from fetal liver, cord blood or adult bone marrow, which is associated with a strongly reduced proliferative potential 26. In addition, an age-dependent increase in the frequency of

putative stem cells, coinciding with impaired functionality and reduced lymphoid potential has also been in the human system 27-29. However, it should be recognized

that essentially all experimental studies have used C57BL/6 mice, and it is clear that aging (also of the hematopoietic system) is at least qualitatively different in distinct strains of mice 30, 31. These genetic disparities are also likely to be very prominent in

human.

Figure 1. Hypothetical tracing of the offspring of a single hematopoietic stem cell during aging. In

mice, the most primitive hematopoietic stem cells are believed to cycle only once every ~ 4 months 23.

With each cell division, daughter cells lose developmental (long-term repopulating) potential, such that each daughter is less potent than its ancestor. Cell cycle times decrease with developmental stage. In young mice, the pool of stem cells is small, but the potency of each stem cell is high. In aged mice, the pool of stem cells has expanded, but their functionality is restricted.

Heterogeneity of HSC aging

While there is general consensus on the functional decline of aged hematopoietic stem cells in the mouse, the molecular mechanisms that contribute to such stem cell aging are less clear, and are at times disputed. Partially, our limited insight into the molecular mechanisms that cause HSC to age may result from the fact that most studies in this field have been performed using populations of -inevitably- heterogeneous HSCs.

(14)

Although flow cytometry has allowed the prospective isolation of single cells, the purity of HSCs, at least when assayed in transplantation experiments, is never more than 50%. Therefore, even after the most stringent enrichment protocols, many non-HSCs (mostly progenitors) remain. This is particularly problematic in the aging field, as there is solid evidence that the functional heterogeneity of the HSC population increases with age. Whereas in young mice many HSCs behave qualitatively similar, in aged mice, that display an overall increase in HSC pool size, individual HSCs behave very differently 10, 13, 17, 32, 33. Therefore, it remains possible that even in aged

mice, there are still substantial numbers of very potent (‘young-like’) HSCs, which are diluted by an expanded pool of less potent aged cells (Figure 2).

In order to provide further insight into the age-dependent increase in heterogeneity of the HSC pool, the identification of cell surface markers that allow to prospectively isolate these different populations will be required. Multiple single cell RNA sequencing studies have been performed to elucidate at the molecular level how

heterogeneity may be explained 34-36. Combinatorial single cell techniques such as

single cell transplants, flow cytometry and single-cell RNA have already begun to and will continue to identify HSC subsets that contribute to aging phenotypes. This strategy will boost research into molecular mechanisms of aging through the ability of identifying candidate marker genes (e.g. receptors) for the isolation and functional characterization of ‘aged’ HSCs.

Figure 2: In aged mice the absolute number of cells with regenerative potential increases, but the extent to which individual aged cells contribute to blood cell production becomes highly variable. The relative frequency of stem cells with high

regenerative potential (indicated in white) compared to cells with low regenerative potential (indicated in grey and black) thus decreases upon aging.

(15)

Cell-intrinsic versus cell-extrinsic mechanisms

A topic of much dispute is whether hematopoietic stem cell aging is caused by cell-intrinsic or rather cell-extrinsic parameters. This is not only of academic importance, but also has major implications for potential future interventions. If stem cell aging were largely intrinsically controlled, this would render putative interventions more cumbersome than if aging were the result of an extrinsic perturbation. Experiments with parabiotic mice have attracted quite some attention (understandably, also from the lay press), as the suggestion was raised that unidentified humoral factors that circulate in the blood of young mice could possibly restore cellular functioning in aged mice. Several studies have reported beneficial, arguably anti-aging, effects of young blood on aged vasculature, brain and muscle 37-41, but no beneficial effects on aged hematopoietic

stem cells have been reported. It is important here to remember that essentially all we know on the phenotype of aged hematopoietic stem cells is based on studies in which old stem cells were transplanted into lethally irradiated young recipients 5, 10-12, 15. The

fact that aged HSCs are functionally impaired compared to their young counterparts upon transplantation in young mice, strongly suggest that hematopoietic stem cell aging manifests largely as a consequence of cell-intrinsic molecular changes. Thus, although cell-intrinsic changes may be -partially- dependent on and initiated by changes that occur in the bone marrow microenvironment, transplanting aged HSCs in a young environment apparently cannot reverse these changes. It is evident that during aging

many cellular and structural changes occur in the bone marrow microenvironment 42

, and it is possible that these, as yet poorly understood cell extrinsic changes, affect stem cell intrinsically. Indeed, it has been demonstrated that young bone marrow cells, when transplanted into aged recipients, engraft worse than when transplanted into young recipients 12, 43. Also hematological malignancies can result from a defective

bone marrow environment 44.

Cell-intrinsic mechanisms that cause HSC aging

Multiple molecular and cell-intrinsic mechanisms have been reported to contribute to the age-associated decline of HSC functioning (partially reviewed in 45). Although

mechanistically it may be feasible, and even useful, to separately discuss these multiple aging pathways, in effect they are likely to be highly interdependent and interconnected. While some of these cell-intrinsic causes of aging are unlikely to be reversible, several others in fact might allow interventions, and thus could potentially be explored for pharmacological targeting (Figure 3).

DNA damage

The most irreversible cause of HSC aging relates to the accumulation of random DNA damage. Mice, or patients for that matter, suffering from mutations in genes encoding for proteins involved in DNA repair display many aspects of premature stem cell

aging 46-48 . In addition, aged HSC accumulate signatures of widespread DNA damage,

including γ-H2AX foci 49. To what extent normal hematopoietic stem cell aging in fact

is caused by genetic damage remains unclear. Conceptually, if indeed hematopoietic stem cells are largely quiescent, and divide only rarely in the lifetime of a mouse, it is

(16)

difficult to understand how accumulation of DNA damage could causally contribute to stem cell dysfunction. However, it is possible that cells immediately downstream of quiescent HSCs, that have been shown to be comprised of mostly of myeloid-biased HSCs 50, are target cells to accumulate DNA damage 51.

Beyond random DNA damage, it has recently been shown that in healthy elderly individuals DNA mutations in specific loci are associated with the establishment of clonal hematopoiesis 52 (further extensively discussed below). A specific kind of

DNA damage is caused by erosion of telomeres. While the involvement of telomere shortening in the functional decline of HSC is particularly evident in human 53, also in

mice that have long telomeres, the inability to maintain telomere length is associated

with severe HSC malfunctioning 54. Although the length of telomeres in HSC can

be increased by enforced overexpression of telomerase, in mice this does not rescue functional impairment 55.

Senescence

In many tissues irreversibly cell cycle-arrested senescent cells accumulate during normal aging 56. Conceptually, senescence of stem cells is a ‘contradictio in terminis’

(as an irreversibly arrested stem cell ceases to be a stem cell). Senescence is believed to be predominantly induced by activation of p16, and indeed expression of p16 is considered to be a marker for the presence of senescent cells. The genetic or pharmacological depletion of senescent cells in mice has been shown to enhance

regenerative potential, and indeed extend lifespan 57. Although expression of p16

in aged primitive hematopoietic stem cells has been contested 58, pharmacological

targeting of senescent bone marrow cells has been shown to have beneficial effects 59.

This suggests that putative senescent cells in the bone marrow may secrete factors that negatively affect HSC potential.

Polarity

It has been reported that the asymmetric distribution of specific proteins, collectively referred to as ‘increased polarity’, is a prominent feature of aged HSCs, whereas in

young HSCs this is much less obvious 33, 60. Unequal distribution of these proteins

is believed to be caused by elevated activity of Cdc42. Interestingly, inhibitors of Cdc42 restore polarity in aged HSCs, and improve HSC functioning after transplant. Although it is not straightforward to understand how an acute reversal of the asymmetric

distribution of proteins can have long-term effects on HSC functioning, this study demonstrates that at least some aspects of the aging process appear to be reversible.

Impaired autophagy and mitochondrial activity

A large fraction of aged HSCs show impaired levels of autophagy 33. Autophagy is

generally associated with recycling of organelles, and impaired autophagy in aged HSCs appears to specifically result in the accumulation of mitochondria, which in turn induces metabolic stress. It has been demonstrated that high levels of reactive oxygen species, generated by mitochondria, accumulate in aged HSCs and compromise their

(17)

functioning 61, 62. In addition, reducing mitochondrial stress in aged HSCs can reverse

loss of stem functioning 10,11. Mitochondrial dysfunctioning caused by accumulating

mDNA mutations has been shown to cause multiple hematopoietic defects that are

typically seen in the elderly, but HSCs themselves appear to relatively resistant 63

. Taken together, cellular metabolism, controlled by mitochondrial status and ROS and mTOR signaling, play an important role in maintaining HSC function throughout life, but the molecular cause of age-dependent metabolic derailment remains unclear. Importantly, however, pharmacological interventions in these signaling pathways are feasible and may be exploited to restore function in aged HSCs.

Figure 3: Cell-intrinsic mechanisms that contribute to hematopoietic stem cell aging. While

some of these molecular events are difficult to revert, other may be amenable to pharmacological interventions and could be exploited to rejuvenate HSCs.

Epigenetic reprogramming

Whereas it is not immediately evident how large-scale random DNA damage would accumulate in aged HSCs, which do barely proliferate, it is not difficult to see how an accumulation of aberrant epigenetic marks could readily but gradually lead to loss of stem cell potential. For a ‘true’ self renewal division to occur, a hematopoietic stem cell must, in the timeframe of a single division, not only copy and distribute its entire genome across two daughter cells, but also the plethora of epigenetic marks that cover each and genomic locus must be faithfully reproduced in at least one of the two daughter cells. It seems highly likely that not all epigenetic moieties that are required to specify stem cell functioning are properly maintained after a stem cell has divided. (Figure 4).

(18)

Figure 4. Repressive (indicated by closed symbols) and activating (indicated by open symbols) epigenetic marks cover all genes and affect transcriptionally status. If an HSC divides, all genomic

and epigenomic information must be properly propagated to daughter cells. If epigenetic marks are lost or gained, as in the example above, genes that should be expressed in HSC (B, C, and G) may become repressed, and conversely, genes that should be repressed (A, D, F, and H) may become activated. As a consequence, functional stem cell activity of daughter cell 1 and 2 may be reduced compared to the HSC from which they derive.

Whereas loss or gain of specific epigenetic marks at defined loci may remain inconsequential, collectively chromatin modifications are very important to maintain transcriptional fidelity. The relevance of these epigenetic marks is best exemplified by the fact that perturbation of a large number of epigenetic writers or erasers severely

affects hematopoietic stem cell function 64-68. The gradual and random erosion of

epigenetic marks as stem cells divide and age, provides a conceptual framework which can explain why -at least in the mouse- stem cells can make only a very limited number of true self renewal divisions. As the erosion of epigenetic marks is likely to be different for each and every individual stem cell, with age increased functional cell-to-cell variability is expected to develop. Indeed, this is exactly what has been observed in experimental studies 10, 32. Such increased functional heterogeneity is likely

the result of increased transcriptional ‘noise’, caused by aberrant epigenetic fidelity. This scenario would also suggest that there are no specific ‘aging’ genes, but rather that aging is caused by the cumulative and combinatorial effect of large collections of stochastically differentially expressed genes. In agreement, several distinct gene

(19)

expression profile studies have not been able to identify many commonly differentially expressed transcripts, and conversely, transcripts that have shown to be affected during aging in one study have often not been confirmed in others 12, 69-73.

The involvement of epigenetic regulation in maintaining proper stem cell transcriptional activity and functioning during aging is also suggested by an increasing number of clinical studies in which mutations in genes encoding for epigenetic enzymes have been found in either elderly people whose hematopoietic system has become oligoclonal, or in patients with myelodysplastic syndromes or acute myeloid leukemia

52. Epigenetic genes that are relatively frequently affected in these conditions include

DNMT3A, EZH2, TET2, and SETDB1 74-77. Although the enzymatic activity of these

proteins is rather well understood (DNA methylation, H3K27 tri-methylation, DNA demethylation and H3K9 tri-methylation, respectively), their effect on stem cell functioning is far from clear. Without going into depth here on the specific molecular pathways in which these epigenetic writers are involved, it appears very likely that mutations in these genes -subtly- alter the epigenetic memory of stem cells which, due to large scale (but again, potentially subtle) transcriptional consequences, increases self-renewal potential of mutant cells. Thus, with time, in the bone marrow cells in which these self-renewal-enhancing mutations have occurred are expected to expand clonally. Indeed, studies in mice have revealed that (enforced) altered expression of wild type or mutant epigenetic writers affects self-renewal of HSCs 64, 68, 78.

Clonal hematopoiesis in human

As referred to earlier, clonal hematopoiesis is a frequent event in elderly people. The first studies that reported on this phenomenon assessed whether X-chromosome inactivation patterns were random or skewed in peripheral blood cells of elderly females 79, 80. Indeed, these early studies suggested that during aging blood cells are

derived from fewer and fewer stem cells. Much more recently these early findings have been confirmed, and significantly extended, by multiple independent large-scale sequencing studies 74-77. In addition, these latter studies have shown that clonal

hematopoiesis in otherwise healthy individuals increases the risk to develop leukemia, and, interestingly, cardiovascular disease, and thus is associated with increased mortality. This raises the question whether clonal hematopoiesis is detrimental, and if so, why? It is important to realize that in many perfectly healthy very aged individuals, prominent clonal hematopoiesis is present without any signs of disease 74, 81. Yet, the

fact that age-dependent clonal hematopoiesis is associated with disease suggests that at least in these individuals there may be a causal relationship.

Whether mutations in epigenetic genes that are believed to cause clonal hematopoiesis are truly oncogenic, remains unclear. It is possible that these mutations in fact merely increase self-renewal of benign hematopoietic stem cells, and thus lead to clonal expansion of healthy stem cells. Mutations that are truly leukemogenic are then more likely to occur in such a pool of actively self-renewing stem cells, and thus when patients present with full-blown disease, leukemic cells display frequently mutations

(20)

in epigenetic genes. In such a scenario, it would probably not be appropriate to refer to these genes as leukemia-initiating events.

The fact that subjects with clonal hematopoiesis are more susceptible to develop non-hematological, most notably cardiovascular, disease, suggests that these mutations also affect the functioning of fully differentiated cells, such as monocytes and macrophages. These end-stage cells have been long known to play a role in vascular remodeling, and an age-dependent impaired functioning could explain their association with cardiovascular problems. In fact, whereas ample attention has focused on the aging of hematopoietic stem cells themselves (including the present manuscript) we know much less about the functioning of fully mature peripheral blood cells that are derived from these aged stem cells. It is well established that aged red cells and aged platelets display impaired functioning 82, but is it also true that red cells and

platelets derived from aged stem cells show loss of functioning? It is intriguing that clonal hematopoiesis has been associated with increased incidence of atherosclerotic cardiovascular disease, suggesting that the function of monocytic cells derived from these aged stem cell clones is impaired 83, 84.

Although it has now been well established that clonal hematopoiesis is relatively commonly seen in elderly individuals, it is important to realize that our general understanding of clonal stem cell contribution in hematopoiesis is very limited. We do not know at present how many stem cells actively contribute to blood cell formation during life. Various models have been proposed 85, ranging from clonal succession (at

any given time a few clones are present which become exhausted and replaced by new clones) 86, clonal stability (many clones stably contribute and do so throughout life) 87

, dynamic repetition (many clones contribute but do so at highly distinct efficiencies)

88, or stochastic behavior (clones contribute randomly, their contribution may vary

dramatically with time, they may become extinct or resurface, without any apparent

pattern) 89. The uncertainty as to how many stem cells contribute to steady state

hematopoiesis during aging results from the fact that historically clonal descent of blood cells has been difficult to trace. For clonal analyses, unique heritable markers must be present which discriminate cells from one clone from the other.

Whereas large-scale sequencing efforts which detect spontaneous mutations and their allele frequencies in humans provided insight into clonal make-up in humans, much more precise measurements have been made in experimental conditions using a variety of transgenic approaches. Initial approaches aimed to specifically provide insight into clonal stem cell contribution involved ex vivo barcoding of purified hematopoietic stem cells with retro- or lentiviral vectors that contain unique DNA tags, which were subsequently transplanted into recipient mice 90-93, reviewed in 85.

Later studies embarked on in vivo DNA barcoding of stem cells, thereby avoiding the transplantation process 94, 95. Most recently, in vivo clonal marking in transgenic mice

and fish has been carried out using fluorescent dyes as tags 96, 97. Potentially as a result

(21)

the consistency among these various studies is limited. Whereas some studies indicate that only a limited number of stem cells robustly contributes to blood cell formation 14, 96 others suggest that hematopoiesis is highly polyclonal 90, 94. Irrespective of how these

differences will ultimately be reconciled, it is evident from many experimental studies that mice in which blood cell production is derived from even a single stem cell, are not necessarily prone to develop leukemia 10, 13, 32. Thus, oligo- or even monoclonality

is in and of itself not a (preclinical) sign of pathology.

Is stem cell aging similar in different tissues?

In the field of stem cell biology comparisons between different stem cell-containing tissues are frequently made, with the assumption that at least some general characteristics of stem cells may be conserved across tissues. Although by definition all tissue-specific stem cells are characterized by their self-renewal potential, it is not at all clear whether stem cells age similarly in distinct tissues. In fact, in comparing intestinal stem cells with those of the hematopoietic system, it appears as if aging is very different in these two tissues. Whereas in the hematopoietic system stem cell turnover is very low, proliferation rates in the intestine are very high 98. Intestinal

stem cells do, as expected, accumulate random DNA damage 99, but they do not show

functional decline during normal aging 100. It will be interesting to assess why rapidly

turning-over intestinal stem cells do not age, and slow-turning over hematopoietic stem cells do. In this respect it is interesting to note that another important difference between blood and gut is the extent to which plasticity occurs in progenitor cells in these respective tissues. While in the intestine progenitors have been shown to be able to revert to a stem cell fate 101, in the hematopoietic system such progenitor-to-stem

cell conversions have never been described, at least not during steady state in vivo hematopoiesis. However, enforced expression of transcription factors or microRNAs has been shown to be able to generate transplantable stem cells from committed cell types 102-104.

Future perspectives

Assuming that in humans hematopoietic stem cell aging is not fundamentally different than in mice, it is interesting to speculate that the age-dependent increase of the stem cell pool size is an important factor in the increase of prevalence of myeloproliferative diseases and leukemia in the elderly. An expansion of the pool of primitive cells would increase the number of target cells for malignant derailment. Alternatively, aged stem cells would intrinsically be more susceptible to malignant transformation, possibly as a result of an altered epigenetic landscape, which may have accumulated as a result of repeated proliferation. Also, there is some evidence (but not a lot) that the same oncogenic mutation results in a more aggressive disease in aged stem cells, compared

to young cells 105. If indeed, as in mice, aged human stem cells undergo very few

divisions, it seems implausible that these primitive cells themselves are prone to accumulate multiple mutations.

(22)

useful? At this point in time the answer to both these questions is negative; there is no reliable assay that can quantify human stem cell aging, and if there were such an assay there are no obvious intervention strategies available. However, our understanding of hematopoietic stem cell aging is increasing very rapidly, and we may in the near future well be able to identify individuals who display enhanced stem cell aging. This may offer opportunities to intervene in the kinetics with which hematopoietic stem cells age. Anti-aging interventions may be aimed to prevent, delay, or, most speculatively, reverse -aspects of- the stem cell aging process. Studies in other tissues have demonstrated

that at least delaying components of the aging process can be a viable strategy 106,

and there is no fundamental reason to believe that the hematopoietic system would be exempt for such approaches. In fact, reversal of some aging phenotypes has been achieved by reprogramming old HSCs to pluripotency, and subsequently generating definitive hematopoiesis from these iPS cells 107. Future interventions may be dietary,

pharmacologically, or eventually, cell therapeutically.

Acknowledgements

Studies in the lab of GdH are supported by the European Union FP7 MARie CuRIe AGEing Network Contract number 316964, the Dutch Cancer Society (RUG 2014-7178 ), The Netherlands Organization for Scientific Research/ZonMW, and by the Mouse Clinic for Cancer and Ageing, funded by a grant from the Netherlands Organization of Scientific Research.

References

1. Ossenkoppele G and Lowenberg B. How I treat the older patient with acute myeloid leukemia. Blood. 2015;125(5):767-774.

2. Lopez-Otin C, Blasco MA, Partridge L, Serrano M, Kroemer G. The hallmarks of aging. Cell. 2013;153(6):1194-1217.

3. Harrison DE, Astle CM, Stone M. Numbers and functions of transplantable primitive immunohematopoietic stem cells. Effects of age. J Immunol. 1989;142(11):3833-3840.

4. Stiehm ER, Roberts RL, Hanley-Lopez J, et al. Bone marrow transplantation in severe combined immunodeficiency from a sibling who had received a paternal bone marrow transplant. N Engl J Med. 1996;335(24):1811-1814.

5. Harrison DE and Astle CM. Loss of stem cell repopulating ability upon transplantation. Effects of donor age, cell number, and transplantation procedure. J Exp Med. 1982;156(6):1767-1779. 6. Hellman S, Botnick LE, Hannon EC,

Vigneulle RM. Proliferative capacity of murine hematopoietic stem cells. Proc Natl Acad Sci U S A. 1978;75(1):490-494.

7. Mauch P, Botnick LE, Hannon EC, Obbagy J, Hellman S. Decline in bone marrow proliferative capacity as a function of age. Blood. 1982;60(1):245-252.

8. Ema H, Sudo K, Seita J, et al. Quantification of self-renewal capacity in single hematopoietic stem cells from normal and Lnk-deficient mice. Dev Cell. 2005;8(6):907-914.

9. Wolf NS, Priestley GV, Averill LE. Depletion of reserve in the hemopoietic system: III. Factors affecting the serial transplantation of bone marrow. Exp Hematol. 1983;11(8):762-771.

10. Dykstra B, Olthof S, Schreuder J, Ritsema M, de Haan G. Clonal analysis reveals multiple functional defects of aged murine hematopoietic stem cells. J Exp Med. 2011;208(13):2691-2703 .

11. Van Zant G, Holland BP, Eldridge PW, Chen JJ. Genotype-restricted growth and aging patterns

(23)

hematopoietic stem cell populations of allophenic mice. J Exp Med. 1990;171(5):1547-1565. 12. Rossi DJ, Bryder D, Zahn JM, et al. Cell

intrinsic alterations underlie hematopoietic stem cell aging. Proc Natl Acad Sci U S A. 2005;102(26):9194-9199.

13. Sudo K, Ema H, Morita Y, Nakauchi H. Age-associated characteristics of murine hematopoietic stem cells. J Exp Med. 2000;192(9):1273-1280. 14. Verovskaya E, Broekhuis MJ, Zwart E, et

al. Heterogeneity of young and aged murine hematopoietic stem cells revealed by quantitative clonal analysis using cellular barcoding. Blood. 2013;122(4):523-532.

15. Morrison SJ, Wandycz AM, Akashi K, Globerson A, Weissman IL. The aging of hematopoietic stem cells. Nat Med. 1996;2(9):1011-1016.

16. Geiger H, True JM, de Haan G, Van Zant G. Age- and stage-specific regulation patterns in the hematopoietic stem cell hierarchy. Blood. 2001;98(10):2966-2972.

17. Beerman I, Bhattacharya D, Zandi S, et al. Functionally distinct hematopoietic stem cells modulate hematopoietic lineage potential during aging by a mechanism of clonal expansion. Proc Natl Acad Sci U S A. 2010;107(12):5465-5470. 18. Van Zant G. Studies of hematopoietic stem

cells spared by 5-fluorouracil. J Exp Med. 1984;159(3):679-690.

19. Bradford GB, Williams B, Rossi R, Bertoncello I. Quiescence, cycling, and turnover in the primitive hematopoietic stem cell compartment. Exp Hematol. 1997;25(5):445-453.

20. Cheshier SH, Morrison SJ, Liao X, Weissman IL. In vivo proliferation and cell cycle kinetics of long-term self-renewing hematopoietic stem cells. Proc Natl Acad Sci U S A. 1999;96(6):3120-3125 21. Dykstra B, Kent D, Bowie M, et al. Long-term propagation of distinct hematopoietic differentiation programs in vivo. Cell Stem Cell. 2007;1(2):218-229.

22. Morita Y, Ema H, Nakauchi H. Heterogeneity and hierarchy within the most primitive hematopoietic stem cell compartment. J Exp Med. 2010;207(6):1173-1182.

23. Bernitz JM, Kim HS, MacArthur B, Sieburg H, Moore K. Hematopoietic Stem Cells Count and Remember Self-Renewal Divisions. Cell. 2016;167(5):1296-1309.e10.

24. Van Zant G, de Haan G, Rich IN. Alternatives to stem cell renewal from a developmental viewpoint. Exp Hematol. 1997;25(3):187-192. 25. Jung JJ, Buisman SC, de Haan G. Do hematopoietic stem cells get old? Leukemia. 2017;31(3):529-531.

26. Vaziri H, Dragowska W, Allsopp RC, Thomas TE, Harley CB, Lansdorp PM. Evidence for a mitotic clock in human hematopoietic stem cells: loss of telomeric DNA with age. Proc Natl Acad Sci U S A. 1994;91(21):9857-9860.

27. Kuranda K, Vargaftig J, de la Rochere P, et al. Age-related changes in human hematopoietic stem/

progenitor cells. Aging Cell. 2011;10(3):542-546 28. Pang WW, Price EA, Sahoo D, et al. Human bone marrow hematopoietic stem cells are increased in frequency and myeloid-biased with age. Proc Natl Acad Sci U S A. 2011;108(50):20012-20017.

29. Rundberg Nilsson A, Soneji S, Adolfsson S, Bryder D, Pronk CJ. Human and Murine Hematopoietic Stem Cell Aging Is Associated with Functional Impairments and Intrinsic Megakaryocytic/Erythroid Bias. PLoS One. 2016;11(7):e0158369.

30. De Haan G and Van Zant G. Genetic analysis of hemopoietic cell cycling in mice suggests its involvement in organismal life span. FASEB J. 1999;13(6):707-713.

31. de Haan G and Van Zant G. Intrinsic and extrinsic control of hemopoietic stem cell numbers: mapping of a stem cell gene. J Exp Med. 1997;186(4):529-536.

32. Cho RH, Sieburg HB, Muller-Sieburg CE. A new mechanism for the aging of hematopoietic

(24)

stem cells: aging changes the clonal composition of the stem cell compartment but not individual stem cells. Blood. 2008;111(12):5553-5561. 33. Ho TT, Warr MR, Adelman ER, et al. Autophagy maintains the metabolism and function of young and old stem cells. Nature. 2017;543(7644):205-210.

34. Wilson NK, Kent DG, Buettner F, et al. Combined Single-Cell Functional and Gene Expression Analysis Resolves Heterogeneity within Stem Cell Populations. Cell Stem Cell. 2015;16(6):712-724.

35. Grover A, Sanjuan-Pla A, Thongjuea S, et al. Single-cell RNA sequencing reveals molecular and functional platelet bias of aged haematopoietic stem cells. Nat Commun. 2016;7:11075.

36. Kowalczyk MS, Tirosh I, Heckl D, et al. Single-cell RNA-seq reveals changes in cell cycle and differentiation programs upon aging of hematopoietic stem cells. Genome Res. 2015;25(12):1860-1872.

37. Mayack SR, Shadrach JL, Kim FS, Wagers AJ. Systemic signals regulate ageing and

rejuvenation of blood stem cell niches. Nature. 2010;463(7280):495-500.

38. Katsimpardi L, Litterman NK, Schein PA, et al. Vascular and neurogenic rejuvenation of the aging mouse brain by young systemic factors. Science. 2014;344(6184):630-634.

39. Conboy MJ, Conboy IM, Rando TA. Heterochronic parabiosis: historical perspective and methodological considerations for studies of aging and longevity. Aging Cell. 2013;12(3):525-530.

40. Villeda SA, Plambeck KE, Middeldorp J, et al. Young blood reverses age-related impairments in cognitive function and synaptic plasticity in mice. Nat Med. 2014;20(6):659-663.

41. Castellano JM, Mosher KI, Abbey RJ, et al. Human umbilical cord plasma proteins revitalize hippocampal function in aged mice. Nature. 2017;544(7651):488-492.

42. Kovtonyuk LV, Fritsch K, Feng X, Manz MG,

Takizawa H. Inflamm-Aging of Hematopoiesis, Hematopoietic Stem Cells, and the Bone Marrow Microenvironment. Front Immunol. 2016;7:502. 43. Ergen AV, Boles NC, Goodell MA. Rantes/ Ccl5 influences hematopoietic stem cell subtypes and causes myeloid skewing. Blood. 2012;119(11):2500-2509.

44. Zambetti NA, Ping Z, Chen S, et al. Mesenchymal Inflammation Drives Genotoxic Stress in Hematopoietic Stem Cells and Predicts Disease Evolution in Human Pre-leukemia. Cell Stem Cell. 2016;19(5):613-627.

45. Geiger H, de Haan G, Florian MC. The ageing haematopoietic stem cell compartment. Nat Rev Immunol. 2013;13(5):376-389.

46. Ju Z, Jiang H, Jaworski M, et al. Telomere dysfunction induces environmental alterations limiting hematopoietic stem cell function and engraftment. Nat Med. 2007;13(6):742-747. 47. Rossi DJ, Bryder D, Seita J, Nussenzweig A, Hoeijmakers J, Weissman IL. Deficiencies in DNA damage repair limit the function of haematopoietic stem cells with age. Nature. 2007;447(7145):725-729.

48. Alter BP, Rosenberg PS, Giri N, Baerlocher GM, Lansdorp PM, Savage SA. Telomere length is associated with disease severity and declines with age in dyskeratosis congenita. Haematologica. 2012;97(3):353-359.

49. Walter D, Lier A, Geiselhart A, et al. Exit from dormancy provokes DNA-damage-induced attrition in haematopoietic stem cells. Nature. 2015;520(7548):549-552.

50. Yamamoto R, Morita Y, Ooehara J, et al. Clonal analysis unveils self-renewing lineage-restricted progenitors generated directly from hematopoietic stem cells. Cell. 2013;154(5):1112-1126.

51. Flach J, Bakker ST, Mohrin M, et al. Replication stress is a potent driver of functional decline in ageing haematopoietic stem cells. Nature. 2014;512(7513):198-202.

(25)

hematopoiesis of indeterminate potential and its distinction from myelodysplastic syndromes. Blood. 2015;126(1):9-16.

53. Aubert G and Lansdorp PM. Telomeres and aging. Physiol Rev. 2008;88(2):557-579.

54. Rudolph KL, Chang S, Lee HW, et al. Longevity, stress response, and cancer in aging telomerase-deficient mice. Cell. 1999;96(5):701-712.

55. Allsopp RC, Morin GB, Horner JW, DePinho R, Harley CB, Weissman IL. Effect of TERT over-expression on the long-term transplantation capacity of hematopoietic stem cells. Nat Med. 2003;9(4):369-371.

56. Childs BG, Durik M, Baker DJ, van Deursen JM. Cellular senescence in aging and age-related disease: from mechanisms to therapy. Nat Med. 2015;21(12):1424-1435.

57. Baker DJ, Childs BG, Durik M, et al. Naturally occurring p16(Ink4a)-positive cells shorten healthy lifespan. Nature. 2016;530(7589):184-189.

58. Attema JL, Pronk CJ, Norddahl GL, Nygren JM, Bryder D. Hematopoietic stem cell ageing is uncoupled from p16 INK4A-mediated senescence. Oncogene. 2009;28(22):2238-2243.

59. Chang J, Wang Y, Shao L, et al. Clearance of senescent cells by ABT263 rejuvenates aged hematopoietic stem cells in mice. Nat Med. 2016;22(1):78-83.

60. Florian MC, Dorr K, Niebel A, et al. Cdc42 activity regulates hematopoietic stem cell aging and rejuvenation. Cell Stem Cell. 2012;10(5):520-530.

61. Ito K, Hirao A, Arai F, et al. Reactive oxygen species act through p38 MAPK to limit the lifespan of hematopoietic stem cells. Nat Med. 2006;12(4):446-451.

62. Mohrin M, Shin J, Liu Y, et al. Stem cell aging. A mitochondrial UPR-mediated metabolic checkpoint regulates hematopoietic stem cell aging. Science. 2015;347(6228):1374-1377.

63. Norddahl GL, Pronk CJ, Wahlestedt M, et al. Accumulating mitochondrial DNA mutations drive premature hematopoietic aging phenotypes distinct from physiological stem cell aging. Cell Stem Cell. 2011;8(5):499-510.

64. Kamminga LM, Bystrykh LV, de Boer A, et al. The Polycomb group gene Ezh2 prevents hematopoietic stem cell exhaustion. Blood. 2006;107(5):2170-2179.

65. Iwama A, Oguro H, Negishi M, et al. Enhanced self-renewal of hematopoietic stem cells mediated by the polycomb gene product Bmi-1. Immunity. 2004;21(6):843-851.

66. Lessard J, Schumacher A, Thorsteinsdottir U, van Lohuizen M, Magnuson T, Sauvageau G. Functional antagonism of the Polycomb-Group genes eed and Bmi1 in hemopoietic cell proliferation. Genes Dev. 1999;13(20):2691-2703 67. Majewski IJ, Ritchie ME, Phipson B, et al. Opposing roles of polycomb repressive complexes in hematopoietic stem and progenitor cells. Blood. 2010;116(5):731-739.

68. Klauke K, Radulovic V, Broekhuis M, et al. Polycomb Cbx family members mediate the balance between haematopoietic stem cell self-renewal and differentiation. Nat Cell Biol. 2013;15(4):353-362.

69. Noda S, Ichikawa H, Miyoshi H.Hematopoietic stem cell aging is associated with functional decline and delayed cell cycle progression. Biochem Biophys Res Commun. 2009;383(2):210-215.

70. Wahlestedt M, Norddahl GL, Sten G, et al. An epigenetic component of hematopoietic stem cell aging amenable to reprogramming into a young state. Blood. 2013;121(21):4257-4264.

71. Beerman I, Bock C, Garrison BS, et al. Proliferation-dependent alterations of the DNA methylation landscape underlie hematopoietic stem cell aging. Cell Stem Cell. 2013;12(4):413-425.

72. Sun D, Luo M, Jeong M, et al. Epigenomic profiling of young and aged HSCs reveals concerted changes during aging that reinforce

(26)

83. Fuster JJ, MacLauchlan S, Zuriaga MA, et al. Clonal hematopoiesis associated with TET2 deficiency accelerates atherosclerosis development in mice. Science. 2017;355(6327):842-847. 84. Jaiswal S, Natarajan P, Silver AJ, et al. Clonal Hematopoiesis and Risk of Atherosclerotic Cardiovascular Disease. N Engl J Med. 2017;377(2):111-121.

85. Bystrykh LV, Verovskaya E, Zwart E, Broekhuis M, de Haan G. Counting stem cells: methodological constraints. Nat Methods. 2012;9(6):567-574.

86. Rosendaal M and Adam J. Haemopoiesis by clonal succession? Blood Cells. 1984;10(2-3):473-485.

87. Prchal JT, Prchal JF, Belickova M, et al. Clonal stability of blood cell lineages indicated by X-chromosomal transcriptional polymorphism. J Exp Med. 1996;183(2):561-567.

88. Takizawa H, Regoes RR, Boddupalli CS, Bonhoeffer S, Manz MG. Dynamic variation in cycling of hematopoietic stem cells in steady state and inflammation. J Exp Med. 2011;208(2):273-284.

89. Roeder I, Kamminga LM, Braesel K, Dontje B, de Haan G, Loeffler M. Competitive clonal hematopoiesis in mouse chimeras explained by a stochastic model of stem cell organization. Blood. 2005;105(2):609-616.

90. Lu R, Neff NF, Quake SR, Weissman IL. Tracking single hematopoietic stem cells in vivo using high-throughput sequencing in conjunction with viral genetic barcoding. Nat Biotechnol. 2011;29(10):928-933.

91. Gerrits A, Dykstra B, Kalmykowa OJ, et al. Cellular barcoding tool for clonal analysis in the hematopoietic system. Blood. 2010;115(13):2610-2618.

92. Naik SH, Perie L, Swart E, et al. Diverse and heritable lineage imprinting of early haematopoietic progenitors. Nature. 2013;496(7444):229-232.

93. Cheung AM, Nguyen LV, Carles A, et al. self-renewal. Cell Stem Cell. 2014;14(5):673-688

73. Chambers SM, Shaw CA, Gatza C, Fisk CJ, Donehower LA, Goodell MA. Aging hematopoietic stem cells decline in function and exhibit epigenetic dysregulation. PLoS Biol. 2007;5(8):e201.

74. Zink F, Stacey SN, Norddahl GL, et al. Clonal hematopoiesis, with and without candidate driver mutations, is common in the elderly. Blood. 2017 75. Xie M, Lu C, Wang J, et al. Age-related mutations associated with clonal hematopoietic expansion and malignancies. Nat Med. 2014;20(12):1472-1478.

76. Jaiswal S, Fontanillas P, Flannick J, et al. Age-related clonal hematopoiesis associated with adverse outcomes. N Engl J Med. 2014;371(26):2488-2498.

77. Genovese G, Kahler AK, Handsaker RE, et al. Clonal hematopoiesis and blood-cancer risk inferred from blood DNA sequence. N Engl J Med. 2014;371(26):2477-2487.

78. Mayle A, Yang L, Rodriguez B, et al. Dnmt3a loss predisposes murine hematopoietic stem cells to malignant transformation. Blood. 2015;125(4):629-638.

79. Busque L, Mio R, Mattioli J, et al. Nonrandom X-inactivation patterns in normal females: lyonization ratios vary with age. Blood. 1996;88(1):59-65.

80. Champion KM, Gilbert JG, Asimakopoulos FA, Hinshelwood S, Green AR. Clonal haemopoiesis in normal elderly women: implications for the myeloproliferative disorders and myelodysplastic syndromes. Br J Haematol. 1997;97(4):920-926. 81. van den Akker EB, Pitts SJ, Deelen J, et al. Uncompromised 10-year survival of oldest old carrying somatic mutations in DNMT3A and TET2. Blood. 2016;127(11):1512-1515.

82. Hoehn RS, Jernigan PL, Chang AL, Edwards MJ, Pritts TA. Molecular mechanisms of erythrocyte aging. Biol Chem. 2015;396(6-7):621-631.

(27)

Analysis of the clonal growth and differentiation dynamics of primitive barcoded human cord blood cells in NSG mice. Blood. 2013;122(18):3129-3137.

94. Sun J, Ramos A, Chapman B, et al. Clonal dynamics of native haematopoiesis. Nature. 2014;514(7522):322-327.

95. Busch K, Klapproth K, Barile M, et al. Fundamental properties of unperturbed haematopoiesis from stem cells in vivo. Nature. 2015;518(7540):542-546.

96. Yu VW, Yusuf RZ, Oki T, et al. Epigenetic Memory Underlies Cell-Autonomous Heterogeneous Behavior of Hematopoietic Stem Cells. Cell. 2016;167(5):1310-1322.e17.

97. Henninger J, Santoso B, Hans S, et al. Clonal fate mapping quantifies the number of haematopoietic stem cells that arise during development. Nat Cell Biol. 2017;19(1):17-27. 98. Clevers H. The intestinal crypt, a prototype stem cell compartment. Cell. 2013;154(2):274-284.

99. Blokzijl F, de Ligt J, Jager M, et al. Tissue-specific mutation accumulation in human adult stem cells during life. Nature. 2016;538(7624):260-264.

100. Barker N, Huch M, Kujala P, et al. Lgr5(+ve) stem cells drive self-renewal in the stomach and build long-lived gastric units in vitro. Cell Stem Cell. 2010;6(1):25-36.

101. Tetteh PW, Basak O, Farin HF, et al. Replacement of Lost Lgr5-Positive Stem Cells through Plasticity of Their Enterocyte-Lineage Daughters. Cell Stem Cell. 2016;18(2):203-213. 102. Riddell J, Gazit R, Garrison BS, et al. Reprogramming committed murine blood cells to induced hematopoietic stem cells with defined factors. Cell. 2014;157(3):549-564.

103. Pereira CF, Chang B, Qiu J, et al. Induction of a hemogenic program in mouse fibroblasts. Cell Stem Cell. 2013;13(2):205-218.

104. Wojtowicz EE, Lechman ER, Hermans KG, et al. Ectopic miR-125a Expression Induces

Long-Term Repopulating Stem Cell Capacity in Mouse and Human Hematopoietic Progenitors. Cell Stem Cell. 2016;19(3):383-396.

105. Signer RA, Montecino-Rodriguez E, Witte ON, McLaughlin J, Dorshkind K. Age-related defects in B lymphopoiesis underlie the myeloid dominance of adult leukemia. Blood. 2007;110(6):1831-1839.

106. Goodell MA and Rando TA. Stem cells and healthy aging. Science. 2015;350(6265):1199-1204.

107. Wahlestedt M, Erlandsson E, Kristiansen T, et al. Clonal reversal of ageing-associated stem cell lineage bias via a pluripotent intermediate. Nat Commun. 2017;8:14533.

(28)
(29)
(30)

CHAPTER 2

Genomic and functional integrity of the

hema-topoietic system requires tolerance of oxidative

DNA lesions

Ana Martín-Pardillos1,*, Anastasia Tsaalbi-Shtylik1,*, Si Chen2, Seka Lazare3, Ronald

P. van Os3, Albertina Dethmers-Ausema3, Nima Borhan Fakouri4, Matthias

Boss-hard5, Rossana Aprigliano5,6, Barbara van Loon5,6, Daniela C. F. Salvatori7, Keiji

Hashimoto8, Celia Dingemanse-van der Spek1, Masaaki Moriya8, Lene Juel

Rasmus-sen4, Gerald de Haan3, Marc H.G.P. Raaijmakers2,† and Niels de Wind1,†

1 Department of Human Genetics, Leiden University Medical Center,

2333ZC Leiden, The Netherlands.

2 Department of Hematology, Erasmus Medical Center Cancer Institute,

Rotterdam, 3000CB, the Netherlands

3 European Research Institute for the Biology of Ageing, University

Medical Centre Groningen, University of Groningen, 9713AV Groningen, The Netherlands.

4 Center for Healthy Aging, Department of Cellular and Molecular

Medicine, University of Copenhagen, Copenhagen, DK-2200 Denmark.

5 Department of Molecular Mechanisms of Disease, University of

Zürich, 8057 Zürich, Switzerland.

6 Department of Cancer Research and Molecular Medicine, Norwegian

University for Science and Technology, 7491 Trondheim, Norway

7 Department of Pathology, Leiden University Medical Center, 2333ZC

Leiden, The Netherlands.

8 Department of Pharmacological Sciences, Stony Brook University

Medical School, Stony Brook, NY 11794-8651, USA.

* These authors contributed equally to this work.

† Co-corresponding authors. NdW serves as lead contact. E-mail: n.de_

wind@lumc.nl. Phone: +31715269627, fax: +31715268284 Exp Hematol. 2017 53:26-30

(31)

Abstract

Endogenous DNA damage is causally associated with the functional decline and transformation of stem cells that characterize ageing. DNA lesions that have escaped DNA repair can induce replication stress and genomic breaks that induce senescence and apoptosis. It is not clear how stem and proliferating cells cope with accumulating endogenous DNA lesions, and how these ultimately affect the physiology of cells and tissues. Here we have addressed these questions by investigating the hematopoietic system of mice deficient for Rev1, a core factor in DNA translesion synthesis (TLS), the post-replicative bypass of damaged nucleotides. Rev1 hematopoietic stem and progenitor cells displayed compromised proliferation, and replication stress that could be rescued with an antioxidant. The additional disruption of Xpc, essential for global-genome nucleotide excision repair (ggNER) of helix-distorting nucleotide lesions, resulted in the perinatal loss of hematopoietic stem cells, progressive loss of bone marrow, and fatal aplastic anemia between 3 and 4 months of age. This was associated with replication stress, genomic breaks, DNA damage signaling, senescence and apoptosis in bone marrow. Surprisingly, the collapse of the Rev1Xpc bone marrow was associated with progressive mitochondrial dysfunction and consequent exacerbation of oxidative stress. These data reveal that, to protect its genomic and functional integrity, the hematopoietic system critically depends on the combined activities of repair and replication of helix-distorting oxidative nucleotide lesions by ggNER and Rev1-dependent TLS, respectively. The error-prone nature of TLS may provide mechanistic understanding of the accumulation of mutations in the hematopoietic system upon ageing.

Introduction

The ageing-associated attrition of proliferating tissues is accompanied by mutagenesis and genomic rearrangements, cellular senescence, and mitochondrial dysfunction,

possibly in response to the accumulation of endogenous DNA damage 1. Each cell

in the body acquires 104-105 endogenous DNA lesions per day 2. Most DNA lesions

are repaired by a network of complementary DNA repair systems, each of which deals with a specific class of lesions, as an integral part of the DNA damage response

[reviewed in 3-5]. The dominant, global-genome nucleotide excision repair (ggNER,

Figure 1A) pathway specifically recognizes and removes helix-distorting endogenous

and exogenous nucleotide lesions 6. ggNER deficiency, as exemplified by mice with

a disruption of the Xpc gene, only causes minor phenotypes when the organism is not exposed to exogenous genotoxic agents. This result suggests that unrepaired endogenous helix-distorting DNA lesions can be tolerated at the genome of proliferating cells 6,7.

Unrepaired nucleotide lesions usually arrest processive replication forks, resulting in lesion-containing single-stranded (ss) DNA tracks. Persistent ssDNA tracks can collapse to cytotoxic and recombinogenic double-strand DNA (dsDNA) breaks. To fill such lesion-containing ssDNA tracts and to enable termination of genomic replication, cells have evolved multiple mechanisms, collectively called DNA damage tolerance

(32)

caused by unreplicated damaged nucleotides. DNA translesion synthesis (TLS) is the major DNA damage tolerance mechanism in mammals. TLS employs specialized DNA polymerases to directly replicate across damaged nucleotides. Since these TLS polymerases frequently misincorporate opposite the lesion, cellular survival by TLS comes at the expense of nucleotide substitution mutagenesis (Figure 1A) 9. The core

TLS polymerase Rev1 inserts cytidines opposite abasic nucleotides and a limited spectrum of nucleotide adducts at the minor groove of the DNA helix 10,11. Additionally,

Rev1 plays an important regulatory role in TLS of helix-distorting nucleotide lesions, of (non-damaged) G-quadruplex structures 12, whereas it also operates in the repair of

interstrand DNA crosslinks 9,13.

The long-lived nature of hematopoietic stem cells (HSCs) makes these cells particularly susceptible to endogenous and exogenous genotoxic insults that can limit their functional capacity and that also induce genomic alterations that predispose to hematopoietic malignancies 14. Therefore, the hematopoietic system provides a paradigm to study the

involvement of endogenous DNA damage in development, maintenance, decay and cancer development of proliferating and differentiating tissues 14-18.

Here we investigated the role of Rev1-dependent TLS in the development and maintenance of the hematopoietic system. We reveal the requirement of TLS opposite unrepaired endogenous helix-distorting DNA lesions for the maintenance of HSPCs. We furthermore demonstrate that ggNER and, to a greater extent, TLS provide independent and complementary mechanisms to protect the hematopoietic system against the detrimental phenotypes caused by endogenous DNA lesions.

Methods in brief

Full methods are included in the Supplemental Information.

Mice and cell lines.

All mouse experiments were approved by the ethical review board of the Institute and the, specific pathogen-free, mice were kept according to FELASA guidelines. Wild type, Rev1, Xpc and Rev1Xpc mouse cohorts were obtained by crossing FVB and C57Bl/6 parents. Equal numbers of hybrid males and females were used for most experiments. Transplantation experiments using Rev1 animals were performed in the C57Bl/6 background. Hairless albino SKH-1 mice were used for measuring sensitivity of the skin to ultraviolet light-induced nucleotide lesions.

Competitive transplantation experiments using Rev1 hematopoietic cells were performed as follows: whole bone marrow or isolated HSCs (alone or, when indicated, together with W41.SJL (c-kit receptor-mutant bone marrow cells), from donors were introduced into lethally irradiated B6.SJL. Secondary transplantation was performed

as previously described 20. Chimerism of the hematopoietic system was analyzed at

different time points after transplantation using multiplex PCR on blood. Reconstitution of the Rev1Xpc bone marrow with Xpc bone marrow was performed at the age of 1.5

(33)

months, by injecting 5x106 Xpc bone marrow cells. Unless stated otherwise mice

were given an intra-peritoneal injection with Bromodeoxyuridine (BrdU) and ethynyl-2’-deoxyuridine (EdU), one hour before killing with CO2, to label replicating cells. MEF lines were obtained by spontaneous immortalization of fibroblasts from 13.5-day embryos of the hybrid background. Survival after genotoxin exposure was measured using clonogenic assays.

Whole blood analysis

For whole blood analysis, peripheral blood was manually and automatically counted. For the quantification of white blood cell ratios and Howell-Jolly bodies, blood smears were stained with Giemsa.

Bone marrow and blood preparation for stainings

Bone marrow cells were extracted by flushing femurs and tibia. One intact femur was used to calculate bone marrow cellularities. Cell numbers were normalized to body weight. Paraffin-embedded sections were stained with Hematoxylin and Eosin (HE). Whole blood was extracted from heart. Erythrocytes were lysed before characterization

of hematopoietic and blood cells.

Analysis of cultured HSCs

Long-term-HSCs were isolated and cultured for two weeks with or without 100µM N-acetylcysteine. Colony sizes were scored at day 7 and day 14. At day 14, cells were collected for immunofluorescence staining of the DNA damage marker γH2AX.

Cobblestone Area-Forming Cell (CAFC) assay

The CAFC assay assesses the clonogenicity and size in vitro of different HSPC populations and was performed as previously described 21,22.

Analysis of HSPC populations

HSPC populations were isolated by fluorescence activated cell sorting (FACS), following labeling of HSPC population-specific surface markers with antibodies, and labeling these with fluorophores. Concentrations and origin of these reagents are depicted in Supplemental Table 1. To identify stromal cells, bone marrow cell suspensions were stained with the stromal cell-specific antibodies followed with fluorophore labeling and FACS. Fetal livers were analyzed after BrdU injection of the mother. Fetal liver cell suspensions were stained with cell surface markers for HSPCs (see above), followed by BrdU staining. For the analysis of proliferation and apoptosis, freshly isolated fetal liver cells were stained with HSPC cell surface markers and then for Ki67 or Annexin V, respectively. 7-Aminoactinomycin D (7-AAD) was added in the cell suspension to stain DNA prior to analysis to exclude dead cells. Data were acquired using flow cytometry.

Immunohistochemistry and immunofluorescence

(34)

for BrdU, Caspase-3, Dec1, γ-H2AX, Ki67, p16, 8OHdG, phospho-p38 or 53BP1. Staining for incorporated EdU was performed according to the manufacturer’s instructions. Staining of bone marrow for 4-hydroxynonenal (4-HNE) was performed on deparaffinized sections after antigen retrieval. Sections were counterstained with Mayers hemalum.

Alkaline comet assays

Alkaline single cell electrophoresis (‘Comet’) assays that enable to detect ssDNA and dsDNA breaks at the genome, and staining for incorporated BrdU to identify S phase nuclei, were performed on bone marrow cell suspensions, essentially as described 23.

TLS assay

The generation of a site-specific single-stranded H-εdC lesion and the determination of the efficiency and mutagenicity of TLS were performed essentially as described for a benzo[a]pyrene-dG adduct 24.

Analysis of mitochondrial function

Mitochondrial membrane potentials were measured by flow cytometry. Measurement of mtDNA was performed by quantitative PCR. The relative mtDNA levels were calculated using formula 2×2ΔCT. Mitochondrial respiration of freshly isolated total bone marrow cells was analyzed using a Seahorse extracellular flux bioanalyzer. For Western blotting of mitochondrial proteins, freshly frozen bone marrow cells were re-suspended in RIPA buffer followed by brief sonication. Electrophoresis, blotting, antibody incubations and visualization of signals using enhanced chemiluminescence was performed using standard protocols.

Results

Rev1 contributes to HSPC maintenance.

Rev1-deficient mice displayed mild cytopenia of the blood, affecting all lineages (Figure 1B, 1C). Analysis of short- and long-term hematopoietic stem cell (HSC) populations revealed a slight, but significant, decrease in the frequency of early hematopoietic progenitors (LSK cells), already at a young age (Figure 1D). Competitive repopulation experiments revealed that long-term Rev1 HSCs (the LSK-SLAM population)

were unable to compete with simultaneously administered W41.SJL HSCs 20 [that

themselves display compromised repopulation capacity] in reconstituting the bone marrow of lethally irradiated wild type mice (Figure 1E). Even in the absence of W41. SJL competitors, Rev1 HSCs repopulated the recipients’ hematopoietic system only inefficiently (Supplemental Figure 1A). This phenotype was further aggravated in serial transplantations, indicating a persistent disadvantage of Rev1 HSCs (Supplemental Figure 1B). HSCs isolated from 14-day-old fetal livers already displayed reduced repopulation capacity (Supplemental Figure 1C), demonstrating that the attenuation of HSC function occurs early in development and independently of the bone marrow environment. The in vitro clonogenicity of Rev1 HSPCs was significantly reduced,

(35)

compared with wild type (Figure 1F), and Rev1 HSC clones were smaller than controls (Figure 1G). The combined defects in transplantability and in vitro growth strongly suggest that Rev1-deficient HSPCs display compromised proliferative capacity, which most likely is cell-intrinsic.

ggNER and TLS synergize to protect the hematopoietic system from the genotoxicity of helix-distorting DNA lesions.

We argued that, in case the proliferative defects of Rev1 HSPCs reflect perturbed TLS of endogenous helix-distorting nucleotide lesions, inactivation of ggNER of this class of nucleotide lesions would exacerbate the hematopoietic phenotypes (Figure 1A). Indeed, disruption of Xpc synergistically increased the sensitivity of both Rev1-deficient skin and cultured embryonic fibroblasts to helix-distorting photolesions

induced by UV light 6 (Supplemental Figures 2A, 2B). Previously we have shown that,

upon UV exposure, Rev1Xpc fibroblasts display high levels of replication stress, caused by replicons arrested at unrepaired helix-distorting photolesions 25. Thus, ggNER and

Rev1-dependent TLS jointly protect cells from the genotoxic effects of UV light by repairing or tolerating photolesions, respectively (see Figure 1A). Rev1Xpc embryos were significantly smaller than wild type or single-deficient littermates, and the double-deficient mice were born at sub-Mendelian ratios (Supplemental Figures 2C, 2D), consistent with enhanced sensitivity to endogenous helix-distorting nucleotide lesions. While Xpc mice displayed near-wild type lifespans, Rev1 single-deficient mice died

slightly earlier than their wild type littermates (20 versus 23 months of age, on the average; Figure 2A) of various causes unrelated to the hematopoietic defects. This shortened life span suggests that ggNER is not sufficient to repair all of its substrates, making proliferating cells dependent on Rev1-mediated TLS. In contrast, TLS+ggNER double-deficient, Rev1Xpc mice died at, on the average, 3.5 months of age (Figure 2A) displaying aplasia of the bone marrow and pancytopenia of bone marrow and blood that most markedly affected neutrophils (Figure 2B, 2C, 2D, Supplemental Table 2, Supplemental Figure 2E-2H). HSC counts were strongly reduced already in the liver of Rev1Xpc fetuses, and this reduction aggravated through life (Figure 2E, Supplemental Figure 2I). In conclusion, the hematopoietic phenotypes of Rev1 deficient mice are exacerbated synergistically by concomitant ggNER deficiency, which provides a strong argument that failure to replicate endogenous helix-distorting nucleotide lesions results in hematopoietic attrition. We then investigated the fate of Rev1Xpc embryonic liver HSCs. All Rev1Xpc HSC subpopulations, and also long-term progenitor cells, were characterized by a reduced G0 and increased Ki67-positive S/G2/M fraction, suggesting exit from quiescence and enhanced proliferation or delay in progression through S phase (Figure 2F, Supplemental Figure 2J). In support with the latter, the fraction of Rev1Xpc LSK cells and short-term HSC that resided in S phase was increased in Rev1Xpc embryonic liver HSC (Figure 2G, Supplemental Figure 2K). The concurrent emergence of a BrdU-positive sub-G0 LKS fraction suggested increased death of proliferating Rev1Xpc HSC (Figure 2G, Supplemental Figure 2K) although we could not detect increased apoptosis (Supplemental Figure 2L).

Referenties

GERELATEERDE DOCUMENTEN

It seems plausible that with each cell division the potential of a hematopoietic stem cell to contribute to blood cell production declines, and that simultaneously the pool of

In live bone marrow cells of 3 months old Rev1Xpc mice the basal, ATP-dependent and reserve respiratory capacities in viable cells had virtually vanished (Figure 5N;

Although caloric restriction prevented age-dependent increase in bone marrow cellularity, caloric restriction was not able to prevent functional decline of aged

Our own analysis allowed a comparison of variability of gene expression between individual mice, in contrast to the 6 previously published studies in which HSCs from multiple

or tissue specific conditional knock-out mouse may be beneficial to investigate at what stages Neogenin expression is upregulated. The possibility of isolating HSCs

The fact that (old) HSCs vary in their gene expression changes may offer some explanation as to the lack of consensus between studies using pooled mice, and highlight an importance

Omdat wij genexpressie in individuele muizen gekarakteriseerd hebben, waren we in staat om retrospectief differentieel tot expressie gebrachte genen te analysern en zien dat

Elucidation of molecular changes of hematopioietic stem cell aging currently lags behind the well-documented phenotypic changes and warrants extensive investigation (This