Platelet Production: Cellular and Molecular Regulation



Platelet Production: Cellular and Molecular Regulation


Elisabeth Cramer BordÉ

William Vainchenker



Platelets are derived from the fragmentation of the cytoplasm of giant bone marrow cells, the megakaryocytes (MKs). MKs are the largest cells of the marrow, up to 50 µm in diameter, and are characterized by a single polylobulated nucleus. They constitute <0.05% of marrow cells. MKs arise and mature in the bone marrow, along with the precursors of other blood cells. Immature MKs are located near the osteoblastic niche and migrate to the subenthothelium of venous sinusoids during maturation (FIGURE 23.1). Given a normal platelet count (150 to 400 × 109/L) and platelet life time (8 to 10 days in humans), an average of 1011 platelets are produced every day to maintain the blood platelet level.

MKs are unique cells that display a single polyploid nucleus and grow by undergoing a unique phenomenon called endomitosis during which the nucleus increases its deoxyribonucleic acid (DNA) content without dividing. Following polyploidization, the cytoplasmic mass of MK increases roughly proportionally to the ploidy status. After achieving full maturation, the MK cytosol is filled with a network of membranes formed by invagination of the plasma membrane, is loaded with secretory granules, and is ready to deliver a myriad of newly formed platelets. Platelets detach from cytoplasmic MK protrusions called proplatelets that extend into the lumen of bone marrow sinusoids and are shed into the peripheral blood from the tips of the extensions. Study of MKs is difficult because of their low numbers in the bone marrow, but scientific advances and techniques such as intravital imaging have aided our understanding of MK maturation and platelet formation. The development of in vitro MK culture systems, allowing the preparation of pure MK populations and the identification of specific and early markers permitting the immunologic detection of MK and their precursors, have made the study of MK feasible. Finally, the identification, purification, and cloning of thrombopoietin (TPO), the major regulator of platelet production, and the availability of recombinant TPO for experimental purposes have facilitated the preparation of large enriched MK populations.

This chapter focuses on the cellular and molecular aspects of MK differentiation and its regulation in humans. Descriptions of relevant animal models are presented as well. TPO and its receptor MPL-R are discussed in detail in Chapter 25 and details of platelet maturation are provided in Chapter 24.


MEGAKARYOCYTE DIFFERENTIATION


The Megakaryocytic Cell

Platelet production ultimately results from the differentiation of a multipotent hematopoietic stem cell (HSC) into MK. HSCs are the only long-lived cells of the hematopoietic system and are capable of regenerating all the hematopoietic tissue by virtue of their self-renewal capability. When HSCs commit to MK differentiation, they lose self-renewal capability and multipotency. Committed cells are called hematopoietic progenitors. The loss of multipotency occurs through precise developmental stages. In a first step, the HSC gives rise to a common myeloid progenitor and a common lymphoid progenitor. The common myeloid progenitors subsequently commit toward specific lineages. The erythroid and MK lineages derive from a common progenitor called either MK/erythrocyte bipotent progenitor (MEP) or BFU-E/MK.1 This bipotent progenitor has been characterized in mice and humans and gives rise to colonies containing mostly erythroblasts and a few MKs. The phenotype of this progenitor has been clearly assessed in the mouse,2 whereas in human, its purification is more preliminary.

At a subsequent developmental stage, the MEP is only committed toward the MK lineage and its following differentiation stages can be divided into three developmental steps. The MK progenitor cell is capable of proliferation, giving rise in vitro to MK colonies. It further differentiates, through a variable number of mitosis, to a transitional cell (promegakaryoblast [PMKB]) that enters an endomitotic process with an average of three DNA cycles of duplication (modal ploidy of 16N). These cells correspond to MK that have begun to synthesize the main platelet proteins and are morphologically identifiable. Cytoplasmic maturation subsequently accelerates to yield a typical MK. Finally, the mature MK sheds platelets in the circulation outside the marrow itself.


Megakaryocytic Progenitors

In the hematopoietic hierarchy, MK progenitors derive from a common MEP, but the MK lineage begins with the appearance of committed MK progenitors. These were first defined by their proliferative capacities in several clonal assays, allowing the identification of three main types of MK progenitors according to colony size and the time required for colonies to develop.3 More recently, immunologic phenotyping has been used for more precise delineation of MK progenitor compartments and better understanding of megakaryopoiesis, especially at the molecular level.

BFU-MK. These are the most primitive MK-committed progenitors, producing colonies composed of more than 50 cells organized into subcolonies. BFU-MK follow the same pattern as BFU-E-derived colonies. They mature in 12 days in mice4 and in 21 days in humans.5

CFU-MK. These cells differ from the BFU-MK by a lower capacity for proliferation. They give rise to colonies containing 3 to 50 cells in 5 days in mice6 and in 12 days in humans.7 Unlike BFU-MK that are in the G0/G1 phase of the cell cycle, CFU-MK are in an active cell cycle.







FIGURE 23.1 Normal human bone marrow biopsy immunostained for fibrinogen. Labeling detects blood vessels and MKs. MKs are frequently located along a vascular sinusoid (V) depicting the vascular niche. Fibrinogen displays a centrifugal staining pattern typical of an α-granule protein. The staining intensity is weak in the small immature MKs (arrow) and it is maximal in large, mature MKs (arrowheads). (Adapted from de Larouziere et al. Inverse immunostaining pattern for synthesized versus endocytosed alpha-granule proteins in human bone marrow megakaryocytes. Br J Haematol 1998;101:618-625, with permission.) Inset: Location of MK in vascular niche. (Adapted from Avecilla et al. Chemokine-mediated interaction of hematopoietic progenitors with the bone marrow vascular niche is required for thrombopoiesis. Nat Med 2004;10:64-71, with permission.)

Late progenitors. In the mouse, these are MK progenitors with a density of <1.050 mg/mL. In 2 to 3 days, they give rise to colonies composed of a few MK with high ploidy.8 The developmental stage of these progenitors is extremely close to that of the transitional cell that has switched toward an endomitotic process.


Megakaryocyte Diploid Precursors

The size of diploid precursors is that of a small lymphocyte and they have no characteristic morphologic features. Thus, their identification depends on the presence of specific immunologic markers.9 Electron microscopy is also a sensitive tool for the cytochemical detection of platelet peroxidase.10

Classification of megakaryocyte maturation stages. MK differentiation is a continuous process, but for the sake of clarity it has been classified into three distinct maturation stages (FIGURE 23.2).11 Megakaryocytic differentiation becomes morphologically identifiable because cell size increases after nuclear ploidy becomes a multiple of 2N. On panoptic staining, the type I or immature MKs (FIGURE 23.2A) have a basophilic cytoplasm and can be distinguished by a large polyploid nucleus, frequent cytoplasm blebs, and relatively large cell volume (i.e., >14 µm in diameter). At this stage, no further cell division occurs and the cells only undergo endomitosis. The nuclear to cytoplasm ratio (N/C) is high, chromatin thin, and cytoplasm basophilic without azurophilic granules. Type II MKs of intermediate maturation (FIGURE 23.2B) are larger (i.e., 15 to 40 µm), with a polylobulated nucleus and a deep blue cytoplasm. The intense basophilia coincides with high RNA content, a large number of free ribosomes, and a well-developed rough endoplasmic reticulum (ER), indicative of active protein synthesis. In some, an azurophilic area is visible near the cell center where granule formation begins. As the cell matures, the N/C ratio decreases while the amount of cytoplasm dramatically increases and becomes azurophilic, the nuclear lobes become more distinct, and the chromatin condenses. The mature MK type III (FIGURE 23.2C) has a low N/C ratio. The nucleus is dark purple, with a dense chromatin texture and clearly distinct nuclear lobes. The cytoplasm is vast, uniformly granular, and azurophilic. At this stage, the MK is ready to fragment and liberate platelets. It has been estimated that each MK can produce 400 to 8,000 platelets.12







FIGURE 23.2 Light microscopic appearance of the three different maturation stages of MKs. A: Immature MK or megakaryoblast (type I): The relatively large size of this otherwise poorly differentiated cell with it high nucleus/cytoplasm ratio, thin chromatin, and basophilic cytoplasm allows it to be assigned to the MK lineage. B: Intermediate MK (type II): Type II cells are characterized by large convoluted polyploid large nuclei surrounded by a uniformly basophilic cytoplasm. Azurophilic granules appear toward the cell center. C: Mature MK (type III): The MK appears as a large cell with a polylobulated nucleus and a uniformly granular azurophilic cytoplasm.


ULTRASTRUCTURE OF MK DIFFERENTIATION

MK differentiation is characterized by structural changes that are easily observed under the electron microscope. Immunoelectron microscopy, in particular immunogold labeling, is a sensitive tool that has greatly contributed to the description of this process.

Demarcation membrane system. During the maturation process, the demarcation membrane system (DMS) develops as a network of smooth membrane channels located initially near the plasma membrane. These channels derive from multiple invaginations of the MK plasma membrane13 (FIGURE 23.3). The demarcation membranes are always in contact with the extracellular medium, as demonstrated by the penetration of electrondense extracellular tracers such as horseradish peroxidase. The DMS grows rapidly and becomes widespread within the whole cytoplasmic volume, expanding by more than 700% within 72 hours. Although some differences between the protein composition of the DMS and plasma membrane have been detected by freeze fracture studies,14 the main platelet surface (glycoprotein [GP]) receptors are expressed equally on both.15







FIGURE 23.3 Electron micrograph of maturing MKs. Mature MKs are characterized by their large size and multilobed nucleus. The cells contain specific organelles (α-granules [A]) and a well-developed demarcation membrane system (dm) regularly scattered throughout the cytoplasm. (Magnification ×5,940.)

The DMS is the source of proplatelet membrane16 and is the precursor of the membranes of future platelets (plasma membrane and surface connected canalicular system); the mechanism of platelet shedding is discussed in detail below.

Because mature MKs accumulate near marrow sinusoids and represent a potential physical barrier for other marrow cells migrating into the circulation, migrating cells are able to penetrate the open channels of the DMS and transmigrate across the MK volume. This natural phenomenon, termed “emperipolesis,” is occasionally observed in normal bone marrow and is distinct from phagocytosis since the emigrating cell maintains its integrity.17

Golgi complexes and centrioles. Several Golgi complexes within a single MK follow the migration of several centrioles. The localization of centrioles within the cell is random, as are the Golgi complexes. The latter are found more frequently near the cell center in the perinuclear region. Their number appears to be roughly proportional to ploidy. Endogenously synthesized platelet granule proteins can be initially detected immunologically in the trans-Golgi network (TGN) where they undergo glycosylation, concentration, and precipitation. Granule proteins endocytosed from the extracellular medium are not detected in the TGN.18 Membrane GPs also transit through the Golgi complex. It is noteworthy that immunolabeling of the GP Ib/IX/V complex is weak, whereas GPIIb/IIIa and P-selectin, major components of the α-granule membrane, are seen more prominently in the Golgi cisterns and arising vesicles.19

Endoplasmic reticulum. Rough ER and free ribosomes are abundant in immature MK. The smooth ER that derives from the rough ER after loss of ribosomes remains abundant until the end of the maturation process and is closely associated with the DMS. It contains a number of enzymes, including NADH-cytochrome c reductase, cyclooxygenase, thromboxane synthase, and Ca2+ Mg2+ adenosine triphosphatase. In humans, a peroxidase activity, the platelet peroxidase, is located in the ER and reflects the presence of cyclooxygenase, an early marker of the cell line. The ER gives rise to the platelet-dense tubular system containing the same cytochemical activity.10 In rodents, acetylcholinesterase (AchE), a marker absent from mature human MK, is also specifically located in the ER.

Formation and packaging of cytosolic granules. Four different types of granules are formed during MK maturation: α-granules, dense granules, lysosomes, and peroxisomes. Their contents are detailed in Table 23.1; the following sections focus on their mechanism of formation.

α-Granules. α-Granules appear early during MK maturation, concurrent with the development of the DMS.20 They are the more numerous granule population in the mature MK (FIGURE 23.3). α-Granules arise from the TGN, where a dark nucleoid is rapidly
visible within budding vesicles. α-Granules are unique organelles in that they acquire their protein contents by two distinct mechanisms: (a) MK biosynthesis21 and (b) plasma protein endocytosis, either receptor mediated or through a fluid phase.22,23 Endocytosis may continue in platelets.24,25 The cytoplasmic markers are initially detected in a diffuse staining pattern when located in the cisterns of synthesis and later in a granular pattern when packaged in granules. Two types of information have been provided by immunolabeling α-granules proteins. First, endocytosed proteins (e.g., fibrinogen and immunoglobulins) appear later in the MK than endogenously synthesized proteins (e.g., von Willebrand factor [vWF]),18 and their distribution patterns within the maturing MK cytoplasm are distinct. Endocytosed proteins are centrifugal and maximal at the cell periphery, whereas synthesized proteins are centripetal and predominant in the juxtanuclear area.26 vWF multimers are apparent in the form of 20-nm tubular structures, first appearing in the TGN and then in the α-granules.27 Stored proteins may not be uniformly distributed among α-granules. For example, angiogenic proteins seem to be selectively located in distinct granule populations according to their function, pro- and antiangiogenic, leading to differential release from platelets.28 The α-granule membrane contains numerous receptors that appear during MK maturation: P-selectin (CD62P), GPIIb/IIIa, CD36, CD9, and PECAM1.19,29 Some are detectable in the Golgi apparatus and appear to be transported directly to α-granules, but others appear to transit via the plasma membrane and to be internalized. Evidence based on intracellular trafficking suggests that the GPIIb/IIIa present in the α-granule membrane originates directly from an internalized plasma membrane pool.30 Finally, fully formed platelet α-granules appear as highly elaborated and compartmentalized structures, storing hemostatic and adhesive proteins and numerous growth factors (Table 23.1). Multivesicular bodies, present in MK and only occasionally in platelets, represent a sorting compartment, intermediately between the TGN and α-and dense granules.31,32 In contrast to secretory cells such as endothelial cells, virtually no constitutive granule secretion occurs in MK; secretion follows the regulated pathway and secreted proteins are packaged within secretory granules.33








Table 23.1 α-Granule proteins

























































Matrix


Membrane


PRESENT IN PLASMA


α-GRANULE MEMBRANE RESTRICTED


(Endocytosed)


P selectin (CD62P)


Fibrinogen


GMP33


Albumin


Osteonectin


Fibronectin


IgG, IgA, IgM (MK synthesized)


α-GRANULE MEMBRANE AND PLASMA MEMBRANE


vWF


GP IIb/IIIa (αIIb/β3)


Coagulation factor V


GP IV


Thrombospondin


CD9


ABSENT FROM PLASMA (Platelet specific)


PECAM1


Multimerin


GP Ib/IX/V


β-Thromboglobulin, PF4,


rap1b


MIXED


VEGF/VPF


PDGF, BFGF EGF, TGF-β


A2-macroglobulin, Matrix Metalloproteinases



Endostatin, angiostatin


Plasminogen activator inhibitor 1



Ig, immunoglobulin; MK, megakaryocyte; vWF, von Willebrand factor; PDGF, platelet-derived growth factor; BFGF, basic fibroblast growth factor; EGF, endothelial growth factor; TGF-β, transforming growth factor-β; GP, glycoprotein; PECAM1, platelet/endothelial cell adhesion molecule-1; VEGF/VPF, vascular endothelial growth factor/vascular permeability factor.


Dense granules. Dense granules are the final marker of the maturation of MK cytoplasm at the ultrastructural level.34 Indeed, these granules acquire their dense appearance quite late, but their limiting membrane and intrinsic receptors are probably formed earlier.35 They arise from TGN and their constituents are sorted from the α-granule components in multivesicular bodies.32 They are lysosome-related organelles containing high concentrations of several low molecular weight molecules necessary for normal blood hemostasis. These include calcium, serotonin, adenine nucleotides, pyrophosphate, and polyphosphate. Their absence results in the hemorrhagic δ-storage pool deficiency disorders. In the Hermansky-Pudlak syndrome, δ-storage pool deficiency is associated with tyrosinase-positive albinism, revealing a relationship between platelet-dense granules and the melanosomes of melanocytes. There are similarities in the biogenesis and secretory code of both types of granules that are linked to the small G-protein rab27 deficient in the ashen strain of mice.36 Murine models have also shown that a specific gene (Slc35d3 gene) is involved in dense granule formation independent of lysosomes and melanosomes.37

Lysosomes and peroxisomes. These granules can be detected cytochemically within maturing MK because of their specific enzymatic content-acid hydrolases38 and catalase, respectively.

The MK nucleus. The nucleus of MK displays several sparse rounded lobes with abundant euchromatin and conspicuous nucleoli (type I) (FIGURE 23.2A). Its maturation is characterized by a constant size increase and segmentation and by chromatin clumping (type II) (FIGURE 23.2B). Through the unique process endomitosis, nuclear lobulation and ploidy increase without division, with the DNA duplicating to levels of 4N to 64N (rarely 128N). A correlation between ploidy and the number of nuclear lobes has been found on squash preparations. Close to the final maturation stage of MK, the nucleus changes texture and aspect (type III) (FIGURE 23.2C). Electron microscopy shows that the nuclear lobes tend to elongate and gather, the chromatin becomes coarser, and long clefts of cytoplasm extend between each lobe. These structural changes herald the start of proplatelet formation and platelet shedding.

Megakaryocyte activation. MKs are able to react to agonists in a manner similar to platelets.39 Thrombin induces dramatic morphologic changes as early as the megakaryoblastic stage, indicating the early expression of protease-activated receptors. In the presence of agonists, the MK plasma membrane becomes bristled and sends out numerous thin pseudopods, the nucleus becomes eccentric, cytosolic organelles centralize, and secretory granules fuse with DMS after which they discharge their contents. This phenomenon, which leads to inappropriate secretion of granule contents, including growth factors, may be implicated in pathologic states such as fibroblast activation and myelofibrosis.40


STAGE-SPECIFIC DIFFERENTIATION MARKERS


Megakaryocyte Progenitors

Monoclonal antibodies against platelet proteins such as GPIIb/IIIa (CD41b), the GPIb complex (GPIbα, GPIbβ, GPIX and GPV, or CD42a, b, c and d), CD9, and CD150 have been used to identify MK progenitors.

In mice, it has been possible to purify MK progenitors close to homogeneity based on the expression of several antigens: Lin c-kit+Sca-1 FcgRlowIL7Rα– Thy1.1 CD34low CD9+CD41+ and more recently, Lin c-kit+Sca-1 CD41+CD150+.41,42

In humans, MK progenitors were initially characterized by the expression of differentiation markers on their surface that are also present on other hematopoietic progenitors, such as CD34, CD31, and CD133. Based on the expression of HLA-DR, it is possible to distinguish between BFU-MK (HLA-DRlow) and CFU-MK (HLA-DRhigh).5 Late MK progenitors can also be isolated in the CD34+CD41+ population which is enriched in CFU-MK.43


Megakaryocyte Precursors

Markers have been identified for later stages of MK differentiation. In the mouse, AchE staining, easily detected with light microscopy, is a convenient marker for investigating murine megakaryopoiesis in vitro.44 In humans, AchE predominates in the erythroid lineage and cannot be used to identify megakaryocytic cells.

Studies in the mouse have shown the presence in the marrow of small AchE-positive cells with 2N ploidy. These cells are the direct precursors of MK and differentiate into polyploid MK in a few days.45 Most of the 2N cells have no proliferative capacity
and undergo cycles of endoreplication. Their number is greatly increased or reduced in induced thrombocytopenia or thrombocytosis, respectively.


Megakaryocytes

Recent studies have used monoclonal antibodies directed against platelet proteins, permitting a better understanding of MK differentiation. Most studies have focused on the main platelet GPs, GPIIb/IIIa (αIIbβ3 or CD41b) and the GPIb complex (GPIbα, GPIbβ, GPIX, and GPV, or CD42a, b, c, and d).

CD41b is composed of CD41a (GPIIb, αIIb) and CD61 (GPIIIa, β3). CD61 is not specific for MK because as the β subunit of the vitronectin receptor avβ3, it can be detected on nonmegakaryocytic cells such as endothelial cells, activated macrophages, and osteoclasts. By contrast, CD41a appears more specific for the MK lineage, especially in adults. In the marrow, a minute cohort of small cells (0.8 to 2.5 × 104 marrow cells) expressing platelet proteins can be detected using an anti-GPIIb antibody.46,47,48 During CFU-MK culture in vitro, a large number of these small cells are present just before the development of recognizable MK colonies (day 6).

CD41 can also be detected on CD34+ cells. Most CD34+ CD41+ cells in the adult correspond to late MK progenitors capable of giving rise to clusters of MK in 5 days43 and to typical CFU-MK. Other experiments in transgenic or knock-in mice, and using the thymidine kinase gene under control of the GPIIb promoter, have shown that after ganciclovir treatment, most hematopoietic progenitors, as well as some HSCs, were killed.49,50

Recently, it has been shown that CD41 is an ontogenic marker of murine hematopoiesis. It is the first marker of hematopoiesis in embryonic stem (ES) cells, as well as in the embryo.51,52 During later development, CD41 becomes restricted to the megakaryocytic lineage, although it is also expressed during mast cell differentiation.51,53 In human neonates, CD41 is not restricted to the megakaryocytic lineage, but it is detected on some hematopoietic progenitors, especially on the bipotent erythroid/MK progenitors, lymphoid progenitors, and some HSCs capable of reconstituting hematopoiesis in immunodeficient animals (SCID-RC).54 It is noteworthy that the expression of Mpl, the TPO receptor, mimics the expression of CD41. It is expressed at a low level on HSCs and all types of hematopoietic progenitors. Its expression markedly increases after commitment toward the MK lineage.55

CD41 expression precedes the detection of other major platelet proteins. The CD42 antigen is detected slightly later on a cell that expresses the CD34 and CD41 antigens. These cells are capable of giving rise to a single MK or a cluster of less than four MKs. The proliferative capacity of these cells is quite similar to those of the human CFU-MK defined in the agar medium.56 CD41+CD42+ cells have been called transitional cells because they can switch from a mitotic stage to an endomitotic stage. In contrast to the CD41, CD42 expression is restricted to the MK lineage, although some late erythroid progenitors may express this marker. To detect the CD42 antigen on the MK surface, the GP Iα (CD42b), GP Iβ (CD42c), and GP IX (CD42a) must be expressed together, whereas GP V (CD42d) is not required and may be synthesized later than the other proteins.57

CD42 expression corresponds to a later differentiation stage that is correlated with the appearance of CD36 (GPIV) and the detection of α-granule proteins such as platelet factor 4 (PF4) and vWF, as well as a marked increase in the expression of Mpl, GPVI, and the integrin α2β1. CD34 disappears during the endomitotic process.

Finally, immature MKs express CD4, which may be relevant for the mechanisms of thrombocytopenia in HIV infection.58,59,60 CD4 expression tends to decrease during MK maturation and is virtually absent from mature MKs and the platelet surface.


ENDOMITOSIS

At the end of the proliferative phase, 2N MK precursors become polyploid. MK polyploidization is unique among mammalian cells in that a multiple of 2N chromosomes is enclosed in a single nucleus surrounded by a single nuclear membrane. To achieve polyploidization, MKs undergo several rounds of DNA duplication and the nucleus segments but without nuclear and cytoplasmic division.

MKs are the only cells that become polyploid during their normal differentiation, whereas other cells may become polyploid after a stress. Polyploidy is a way to increase platelet production because the MK cytoplasm volume increases in parallel with ploidy. An increase in MK ploidy is one of the first events to occur after the induction of acute severe thrombocytopenia in mice: within 24 hours, the modal MK ploidy increases from 16N to 32N. This increase is maximal at 48 hours and ploidy returns to basal values within 120 hours.

As a first maturation step, MKs increase their ploidy and stop DNA duplication at any point between 2N and 64N, rarely going up to 128N. In humans, as in most mammals, the modal ploidy is 16N. However, in some mouse species such as C3H mice, the modal MK ploidy is 32N.61 In the fetus and embryo, MKs have a lower ploidy; most fetal MKs are only 2N or 4N in culture.62 Nonetheless, low ploidy MKs, either on a physiologic basis or due to malignancy, are able to shed platelets in vitro.

MK endomitosis is a rare event that remains difficult to observe. Originally, the term endomitosis was given to the mechanism of polyploidization because it was suggested that the mitotic process was occurring without rupture of the nuclear envelope. However, ultrastructural studies have allowed a description of the endomitotic process. The first steps are identical to those of mitosis. They begin with chromosome condensation and disappearance of the nuclear membrane. Several mitotic spindles are subsequently formed with two centrosomes for a tetraploid nucleus. Each centrosome is active and participates in elongation of the mitotic spindle. At metaphase, chromosomes are aligned at the equator and have sizes similar to those of mitotic cells.

MK culture in the presence of TPO, in conjunction with immunolabeling of various components of the mitotic apparatus, has made it easier to precisely describe the endomitotic process. It was initially thought that endomitosis represented mitosis that had skipped anaphase B and cytokinesis.63,64 In fact, the initiation of endomitosis is identical to a normal mitosis with duplication of the centrosomes, development of a mitotic spindle, chromosome condensation (prophase), rupture of the nuclear envelope, chromosomal alignment at the equatorial plate (metaphase), and separation of the sister chromatids (anaphase).65 Initially, the late phases of endomitosis were not observed because of their rarity. However, with the use of timelapse videomicroscopy, it has been observed that endomitotic
MKs go through telophase and begin a cytokinesis that fails.66,67 At the 2N and 4N transition, the two daughter cells that are nearly fully separated make a backward movement and fuse together. The structure of the central spindle has been a matter of debate, but it appears normal at the 2N and 4N transition. The main abnormality involves the cleavage furrow and its ingression which is incomplete. At higher ploidy, the abnormality is more pronounced with only incomplete cleavage furrow ingression.66,67,68,69 When the MK is polyploid, its spindle is multipolar with the number of poles corresponding to the ploidy level; however, the number of chromosomes at each pole does not correspond to 2N because segregation of chromatids is asymmetrical during endomitosis65 (FIGURE 23.4). It is surprising that a defect in cytokinesis leads to mononucleated cells because it usually induces multinucleated cells. This suggests that anomalies in chromatid separation may be also present during endomitosis. Preliminary evidence suggests that a large fraction of 4N MKs have two nuclei as a consequence of a pure defect in late cytokinesis. Because the abnormality is more pronounced in MK with higher ploidy, the chromatids are less separated, leading to a defect in karyokinesis and to a single nucleus. However, it has recently been shown that polyploid MKs are able to undergo a mitotic cycle with the generation of 2N cells.70






FIGURE 23.4 Immunofluorescent labeling of α and β tubulin in MKs undergoing endomitosis. These images of prometaphases of endomitosis from 4N to 8N, 8N to 16N, and 16N to 32N show the structure and asymmetry of the multipolar mitotic spindles.

The cell cycle of a MK undergoing polyploidization is very similar to a mitotic process. There is clearly a succession of G1/S/G2/M phases, but the M phase is incomplete. After the M phase, the cell reenters G1 to undergo a subsequent cell cycle of DNA duplication. However, the molecular mechanisms responsible for the switch from a mitotic to an endomitotic cycle, as well as the mechanisms involved in stopping the process, are still incompletely understood.


REGULATION OF ENDOMITOSIS

The usual regulators of mitosis are active during the endomitotic process. Cyclin D3, cyclin E, and cyclin A are normally expressed during G, S, and G2/M phases, respectively.71 The presence of cyclin B1 during endomitosis has been a subject of debate. There is now a large body of evidence indicating that cyclin B1 is expressed during the endomitotic cycle and is degraded in anaphase.64,65 The presence of cyclin B1 is absolutely required for an entry in mitosis because the cdc2 (CDK1) kinase activity is indispensable for rupture of the nuclear envelope and chromatid condensation. Cyclin E may also play an important role in ploidization. Cyclin E−/− mice have a marked defect in MK ploidization,72 but a proliferation defect was not observed in the other mitotic cells. In addition, cyclin E overexpression increases MK polyploidy.73

The anaphase-promoting complex is functional in MKs. This proteasome complex regulates the end of mitosis by degrading successive mitotic regulators such as the p55cdc20 subunit and cdh1. During the endomitotic process, cyclin B1, p55cdc20, and Aurora B are normally degraded.65,74 On the other hand, CDK inhibitors such as p21 and p27 are expressed at high levels in polyploid MKs, but are not directly involved in endomitosis. These inhibitors do not seem to contribute to mitotic arrest as MK ploidy is normal in p21−/−, p27−/− and double KO mice.75 However, p19Ink4d appears to play a role in endomitotic arrest because p19Ink4d KO mice have MK with a 64N modal ploidy.76 It is noteworthy that p19Ink4d is transcriptionally regulated by RUNX1 as well as several genes coding for platelet function. Thus, RUNX1 may be a link between the regulation of the end of endomitosis and terminal cytoplasmic differentiation.

There are also differences between endomitosis and mitosis. Levels of cyclin D3 and cyclin D1 are extremely high in MK,71,77 and an increase in cyclin D increases MK ploidy. Cyclin D1 expression is regulated by GATA1 in MK.77 In addition, it has been shown that STAT1, which is downstream of GATA1, also increases polyploidy and MK maturation.78 Interferon γ, which induces STAT1 phosphorylation, induces MK polyploidization as well.

Whatever the molecular mechanisms involved, there are two main abnormalities in the cell cycle of the endomitotic MKs. One abnormality occurs at the end of mitosis and is responsible for the failure in cytokinesis. It has been suggested that this is related to a defect in Aurora B and survivin expression in polyploid MK,79,80 but no major defect in the expression of these proteins and in their
localization in the central spindle has been found at the 2N to 4N transition.74,81 Such abnormalities may be present in higher ploidy MKs.80 Perhaps more germaine to the process, there is accumulation of myosin II at the contractile ring.74 Myosin II normally plays a central role in cytokinesis by promoting the contractile ring forces required for cytokinesis. It is presently unknown if the defect in myosin II accumulation is primary or is secondary to low Rho/Rock pathway activation. The second cell-cycle abnormality occurs at the G1/S transition. A normal cell cannot become polyploid because an as yet unknown “sensor” detects the level of ploidy and blocks subsequent reentry into the S phase. This checkpoint is mainly regulated by p53.82 However, a role for p53 in polyploidization remains controversial.83

Whether polyploidization is advantageous for platelet formation has been unclear. Theoretically, an endomitotic cycle could be shorter than a mitotic cycle because the late phases of the cell cycle are abrogated. However, the time for DNA duplication seems identical in both processes. During terminal differentiation, MK must synthesize large amounts of membrane to produce platelets. The endomitotic process could produce considerable economy of nuclear and plasma membrane synthesis which may aid in the efficiency of platelet production. Polyploidization may also be a way to markedly increase protein synthesis. In favor of this hypothesis, it has been shown that all 2N alleles of a gene remain functional during polyploidization; therefore, polyploidization represents true gene amplification.84 Whether platelets arising from MK of different ploidy have similar or different sizes and function remains a matter of debate. In addition, polyploidy may induce resistance to mutations leading to haploinsufficiency as multiple gene copies are present. Whether polyploidization modifies protein synthesis and gene expression also remains to be determined. It has been reported that there is a linear correlation between mRNA or protein expression and ploidy. However, at high ploidy, expression may reach a plateau. Gene profiling during polyploidization has shown that the processes of polyploidization and terminal differentiation are partially intermingled because expression of genes involved in platelet function increases with ploidy.85 Perhaps this is because terminal differentiation and polyploidization are regulated by the same transcription factors, for example, GATA1, FLI-1, or RUNX1.76,77


PLATELET FORMATION AND SHEDDING

Platelet formation has been described in vitro86 and recently proplatelet formation was observed directly in mice in vivo through vital imaging.87 In vitro, bone marrow MKs produce long extensions of cytoplasm called proplatelets (FIGURE 23.5) giving the MKs, in the process of shedding platelets, the morphologic appearance of an octopus.88

TPO-containing cultures have enabled the study of platelet formation and shedding in vitro using human89 and murine precursors.90 In addition, electron microscopy has enabled the identification of sequential subcellular events that lead to platelet production.89,91 First, there is peripheral redistribution, alignment, and dilatation of demarcation membranes. This is followed by extension of a proplatelet containing an axial bundle of microtubules and regular constriction zones. Finally, there is breakage of a distal constriction rich in microtubules and liberation of a newly formed platelet.






FIGURE 23.5 Photomicrograph of platelet formation by human MKs in culture. After reaching full maturation, the MK cytoplasm is remodeled into long processes termed proplatelets. The proplatelets develop constrictions along their length, giving them a beaded appearance. Finally, there is breakage of distal constrictions rich in microtubules and liberation of newly formed platelets.

Proplatelets bend, branch, and bend again, each elbow and tip being surrounded by a circular microtubule belt.90 Future platelets detach from these cytoplasmic extensions. Alternatively, proplatelets may detach from the cell core and deliver platelets. The extension of proplatelets is accompanied by trafficking of organelles, mainly granules and mitochondria, in both directions, until they reach the tip of the proplatelet where they are immobilized. Videomicroscopy shows that proplatelets are dynamic and reversible structures and that future platelet organelles move freely along their long axes. It could be the circular microtubule at the proplatelet tip that prevents the organelles from going backward again.

Only gold members can continue reading. Log In or Register to continue

Stay updated, free articles. Join our Telegram channel

Jun 21, 2016 | Posted by in HEMATOLOGY | Comments Off on Platelet Production: Cellular and Molecular Regulation

Full access? Get Clinical Tree

Get Clinical Tree app for offline access