Cell organelles The general architecture of an eukaryotic cell (karyon in Greek means nucleus) is quite complex (Fig. 1): the cytoplasm, filling the space between the nucleus and cell membrane, contains a variety of organelles, many of them membranedelimited. Their density and degree of development varies according to the specialization and functional processes occurring in that particular cell. In addition, various filamentous structures form the cell cytoskeleton, where most of these organelles are attached. Fig. 1 General architecture of an eukaryotic cell, showing the membranedelimited and filamentous organelles. Mitochondria (Gr. mitos = thread) are elongated organelles covered by a double membrane bilayer. The inner membrane is folded to the interior, generating a number of protrusions called cristae, which ensure a large surface area. The inner volume of a mitochondrion, called matrix, contains a number of dense granules, representing a store of intracellular calcium. It also contains mitochondrial DNA and RNA, encoding a few dozens of mitochondrial proteins, which prove that mitochondria were originally independent organisms that have chosen at a certain moment during evolution to live in symbiosis with other cells, being incorporated and providing a vital function: completion of aerobic intermediate metabolism of glucose, fatty acids and some aminoacids and their coupling to the respiratory chain and oxidative phosphorylation. Thus, mitochondria function as a main power plant of the cell, generating the largest amount of ATP, the “energy currency” in living systems. Fig. 2 Reactions of the Krebs cycle and their coupling via coenzymes NAD+ and FAD to the respiratory electron transport chain, located in the inner mitochondrial membrane. The proton gradient created across this membrane drives ATP synthesis. The Krebs-Henseleit or tricarboxylic acid cycle (Fig. 2) accomplishes the final degradation of glucose and fatty acids. The end product of anaerobic glycolysis, pyruvate, enters the mitochondrion, where it is converted to acetyl-CoA (coenzyme A), thus beginning the cycle by coupling two carbon units to oxaloacetate and generating citrate. In cells where fatty acid degradation represents a second energy source, like cardiomyocytes, these fatty acids are processed in the mitochondrion as well, by progressive detachment of two carbon units via acetyl-CoA in the Lynen spiral. Degradation of each pyruvate molecule in the Krebs cycle produces 4 reduced nicotinamide-adeninedinucleotide molecules (NADH + H+) and one reduced flavinamide-adeninedinucleotide molecule (FADH2). These oxido-reduction coenzymes are later reoxidized in the respiratory chain, located in the inner mitochondrial membrane (for this reason, the protein content of this membrane is unusually high, approximately 70 %). The respiratory chain (Fig. 2) is composed of a series of metaloproteins called cytochromes, structurally related to hemoglobin, the oxygen carrier. Their cofactor is often heme. This similarity explains the toxic effects of cyanides, carbon monoxide, azides, etc., which complex the heme iron. Other compounds like rotenone and barbiturates inhibit coenzyme Q (ubiquinone). These oxidoreduction enzymes are grouped in 4 complexes (I – IV), and they achieve the task of transferring electrons from a high to a low energy level. During these electron transfers, protons are released in the outer mitochondrial space at well-defined points, as shown in Fig. 2. Thus, a proton gradient is achieved across the inner mitochondrial membrane. According to the chemiosmotic theory of Mitchell (1961), the energy of this proton gradient is converted into ATP by a special proton pump, the f1/f0 ATPase, located in the inner mitochondrial membrane as well. This pump functions on the reverse cycle, i.e. instead of hydrolyzing ATP to extrude protons, it uses proton gradient to synthesize ATP. One ATP molecule is produced for each proton that reenters the matrix. Compounds that increase proton permeability of the inner mitochondrial membrane, like 2,4-dinitrophenol, cancel this proton gradient and uncouple electron transport from oxidative phosphorylation, increasing the production of free oxygen radicals, like the superoxide ion, O22-, an extremely reactive and therefore toxic molecular species. To avoid free radical overproduction, electron transport and oxidative phosphorylation have to be perfectly coupled, producing ATP according to the energy requirements of the cell. Mitochondria are attached to the microtubule network and mobile. They concentrate in areas of increased energy consumption, like the baso-lateral end of tubular renal cells, where they provide ATP for active Na+ transport. Other metabolically active cells, and therefore rich in mitochondria, are neurons, hepatocytes, skeletal muscle fibers and cardiomyocytes. Mitochondria can divide longitudinally, like any bacteria, adapting their density to ATP requirements. During mitotic cell division, each daughter cell receives approximately half of the mitochondria within the parental cell. During fecundation, the egg cell (zygote) keeps the maternal mitochondria within the oocyte. Rybosomes are tiny dense granules, first described using electron microscopy techniques in the years 1950 by George Emil Palade (they were formerly called Palade’s granules). With the panoptic staining method (May-Grünwald – Giemsa or Wright), used on peripheral blood or bone marrow smears, rybosomes appear basophilic. Therefore, cells with increased protein synthesis, like B lymphocytes, which secrete antibodies, or basophilic erythroblasts, which synthesize hemoglobin, have a light-blue appearance of the cytoplasm, resembling a clear, cloudless sky. The detailed structure of ribosomal subunits has been recently revealed using Xray diffraction techniques. It is composed mainly of ribosomal RNA (rRNA), with several enzymes attached in appropriate positions. The large ribosomal subunit features two sites that can adapt transfer RNAs (tRNA). Each tRNA binds specifically an aminoacyl residue at one end, and on a distant loop presents a triplet of nucleotides called anticodon, specified according to the genetic code, identical to the triplet encoding an aminoacyl residue on the leading (3’-5’) DNA strand in the nucleus, and complementary to the triplet on messenger RNA (mRNA). As shown in Fig. 3, mRNA attaches at a specific sequence on the small ribosomal subunit, and thus the translation process begins. The first tRNA attaches to the peptidyl locus, the second on the aminoacyl locus, then specific enzymes detach the first residue and bind it to the second. Further, the small subunit advances like the chariot of a typewriter, carrying the mRNA and the two tRNAs attached to it. The free tRNA of the first residue detaches, and a new tRNA, that of the third residue, binds on the free aminoacyl locus, near the second tRNA having attached a dipeptide. Thus, the process goes on and on, and the polypeptide chain elongates, until the entire gene sequence is translated. Fig. 3 The process of mRNA sequence translation into a polypeptide chain sequence is carried out by ribosomes, composed of two rRNA subunits. The large subunit features two tRNA binding sites: a peptidyl and an aminoacyl site. Rough endoplasmic reticulum (RER) is built of flattened membrane-delimited cisternae communicating between them and with the nuclear envelope. They have a high density of ribosomes attached on their outer surface, and carry out posttranslational processing of export and transmembrane proteins. It is well known that the polypetide sequence of many export proteins begins with a short hydrophobic leader signal sequence (LSS), which is later detached by appropriate proteases. This sequence penetrates the membrane of the rough endoplasmic reticulum and triggers attachement of the ribosome to a polypeptide chain transport protein called translocon. Translocons feature a large hydrophilic channel that allows passage of the elongating polypeptide chain. When new hydrophobic regions of the chain are synthesized (the transmembrane helices of integral proteins), translocons temporarily detach, allowing their insertion into the bilayer, then reattach when the freshly translated sequence becomes again hydrophylic. Thus, several transmembrane-spanning (TM) domains can develop, linked in a specific topology by cytoplasmic and intra-cisternae loops. Post-translational processing of integral and export proteins begins in the RER, as the chain elongates, and continues in the Golgi apparatus. It comprises the extremely important process of folding, which renders the adequate threedimensional shape of the mature protein. A class of specialized proteins called chaperonins (formerly known as heat shock proteins – hsp) achieve a correct folding. The 3D structure is consolidated by disulphide bridges between distant cysteine residues, created by disulphidisomerase. N and C glycosylation follows in the cisternae, via a tunicamycin-inhibited pathway. Lipidation is another important process, helping to anchor membrane proteins to lipid rafts. Several brief sequences on either the cytyoplasmic or the cisternal side of a newly synthesized protein function as retention signals – just a particular case of trafficking signals. For example, the inward rectifier K+ channel subunit KIR6 contains a small C-terminal sequence, RKR, which triggers RER retention. When this sequence is masked by another type of subunit, the sulphonylureea receptor (SUR), the assembly can be trafficked in order to reach cell surface. Other trafficking signals are KKXX (where X signifies any residue) on the cytoplasmic C terminus, or KDEL on the lumen side. Smooth endoplasmic reticulum does not contain ribosomes attached on its surface. It is a branching network of tubules, involved in calcium storage, cholesterol and steroid synthesis, and detoxifying processes. IP3 receptors located on its surface can respond to soluble IP 3 released at the inner cell membrane surface by receptor-mediated PIP2 hydrolysis, and release Ca2+ in the cytoplasm, contributing to important intracellular signaling pathways. Ca 2+ ions are later recaptured in the endoplasmic reticulum via active, ATP-driven transport, by SERCA pumps. Steroid hormones synthesis occurs in specialized endocrine cells, like those of the cortical adrenal gland, ovaris and testis. Drug detoxification occurs predominantly in liver and kidney. For example, in cronic alcohol consumption smooth endoplasmic reticulum becomes abundant in hepatocytes. A specific detoxifying system present in hepatic smooth ER involves cytochrome P450 (maximal absorption wavelength 450 nm). The Golgi apparatus continues trafficking and export of proteins synthesized in the rough endoplasmic reticulum. Transport vesicles secreted by the RER reach the inner (cys or receiving) face of the Golgi apparatus. Additional glycosylation, lipidation and sorting processes occur in the Golgi, then export vesicles are secreted on the outer (trans or shipping) face. There are three types of vesicles secreted by the Golgi apparatus, according to their fate (Fig. 4): 1. vesicles containing proteins to be secreted in the extracellular space; 2. vesicles containing plasma membrane-forming phospholipids and proteins; 3. vesicles containing acid proteases which become lysosomes. Fig. 4 Export protein synthesis and processing begins in the rough ER and continues in the Golgi apparatus, which secretes 3 types of vesicles containing export proteins, membrane components, or acid proteases. Lysosomes contain proteolytic enzymes like acid hydrolases, aryl-sulphatases, 5’-nucleotidases, activated at a pH around 5. They fuse with phagocytosis vesicles or phagosomes forming phagolysosomes. In certain immune system cells, circulating monocytes or tissular macrophages, incomplete digestion of foreign bodies is followed by exposure on cell surface of small fragments called antigens, which trigger cellular or secretory immune responses via T and B lymphocytes. The rich lysosomal contents of monocytes gives their cytoplasm a grey colour at the panoptic staining, like a cloudy sky before a storm. Peroxysomes are distinct membrane-delimited vesicles that achieve free oxygen radicals scavenging. They contain superoxidedismutase and catalase. The cell cytoskeleton is composed of three types of filaments: microtubules, microfilaments, and intermediate filaments (Fig. 5). Fig. 5 Three types of cytoskeletal filaments, their subunit architecture (upper row) and distribution within a cell, as revealed by immunofluorescent staining. Microtubules are tubular structures, 25 nm in diameter, composed of and tubulin subunits.The microtubule system of a cell radiates from a central structure called centrosome or cell center, containing a pair of centrioles, which are symmetrical assemblies of microtubules (9 triads of microtubules around a central core). Mitochondria, lysosomes, and secretory granules move along microtubules using ATP-driven motor molecules (kinesin, dynein, etc.). Microtubules assemble and disassemble continuously. Microfilaments are composed of actin. Monomeric G (globular) actin polymerizes into double helical F (filamentous) actin. This process is regulated by a lot of factors. Microfilaments give the shape of the cell, and are involved in active movements, using myosin molecules as engines. Intermediate filaments are tissue specific: e.g. neurons contain neurofilaments, gliasl cells glial filaments, stratified epithelia of the skin keratin filaments, etc. They are less dynamic than microtubules and microfilaments. Generally they strengthen cells. They form intercellular adhesion elements called desmosomes. Detection of neural or glial intermediate filaments in the amniotic fluid by immunofluorescent monoclonal antibodies can help early diagnosis of neural tube defects. The nucleus is the largest organelle of a cell. It stores the genetic information represented by specific sequences of nucleotides within the non-repeating regions called genes. Usually a gene specifies a single protein sequence. The complete collection of genes of a species represents the genome. Prokaryotic organisms (i.e. those lacking a nucleus, like bacteria or archaea) have a single circular double-stranded DNA comprised of a few dozens of genes, while evolved organisms show much more complex genomes (e.g. the human genome contains an estimated 30,000 to 50,000 genes), arranged in several stretches of DNA called chromosomes. The useful information within a gene is splitted in several coding regions or exons, alternating with non-coding regions or introns. According to the classical dogma of genetics, there is a unique sense of circulation for genetic information: from DNA to messenger RNA (a process called transcription), followed by RNA processing via removal of introns and endto-end joining of exons (splicing), and then by translation from mRNA to protein sequence, according to the genetic code, by free cytoplasmic or RER-attached ribosomes. In addition, chromosomal DNA replicates itself during cell division. Most cells possess a single nucleus, although there are also multinucleate cells, like skeletal muscle fibers, or anucleate cells, like erythrocytes. In order to protect its valuable contents, the nucleus is surrounded by a nuclear envelope, composed of a double bilayer, delimiting a narrow space – the perinuclear cisterna, in continuity with the rough endoplasmic reticulum. Like RER, the nuclear envelope has ribosomes attached to its outer membrane. Nucleotides required for nucleic acid synthesis, as well as freshly synthetized mRNA, enter or leave the nucleus via relatively large nuclear pores. In either optical (e.g. using the classical Giemsa stain) or electron microscopy, the nuclear contents appears as dense regions (heterochromatin) alternating with sparse regions (euchromatin). Heterochromatin contains highly packed DNA, not involved in active transcription processes, while euchromatin, abundant in metabolically active cells, with high protein synthesis rates, contains unfolded DNA representing the genes to be transcribed. Such active cells, like hepatocytes or neurons, feature also one or several nucleoli inside the nucleus. The nucleolus is the site of ribosomal RNA synthesis and assembly. Several regions can be distinguished inside the nucleolus: pars amorpha, pars fibrosa, pars granulosa, pars chromosoma. The individual nuclear DNA stretches can be clearly distinguished only during cell division, when they form highly packed assemblies called chromosomes. Each chromosome contains two identical copies of the double-stranded DNA, packed within the short and long arms (dubbed in genetics p and q). These two copies are united in the middle region named kinetocore, which plays an important role during cell division. Fig. 6 displays the various levels of DNA packing within heterochromatin and chromosomes. The 2-nm diameter DNA double helix is coiled around assemblies formed of eight monomers of specific basic proteins called histones: two subunits of histones H2A, H2B, H3 and H4 – a structure called nucleosome. Another histone subunit, H1, lies between two successive nucleosomes. The next level of twisting is the tight helical fiber (30 nm in diameter), followed by a supercoiled structure – the 200-nm fiber. This fiber forms multiple loops giving rise to chromatids (700-nm fibers) that represent bands within a chromosome. They give specific alternating patterns, which can be evidenced with various stains (Giemsa, acridin orange, etc.). The life cycle of a cell comprises two distinct periods: an interphase, the time interval between two divisions, and the division period called mitotic phase or mitosis (M). Interphase is in turn divided into several periods. G1 (growth 1) follows immediately after a cell division. Active metabolic processes leading to cell growth occur during this time. It is followed by S (synthesis), when DNA within the nucleus is subjected to semiconservative replication: the double helix unfolds, and each strand serves as template for the synthesis of a complementary new strand. Therefore, at the end of the S period, the cell contains a double amount of double helical DNA. During the final growth period (G2) final preparations for cell division are performed. The centrosome completes its division into two identical structures, each composed of a pair of centrioles. The M phase is in turn divided into several periods: prophase, Fig. 6 Various levels of DNA folding within a metaphase, anaphase, chromosome. telophase, and cytokinesis, which can be distinguished with current microscopy techniques. During prophase several events occur in the nucleus and cytoplasm. Nuclear chromatin condenses into diploid chromosomes, each composed of two short and long arms united at the kinetocore – therefore they have the shape of an H or X. The number of chromosomes is speciesspecific: for humans there are 23 pairs of chromosomes (one from the mother and one from the father in each pair), 22 pairs of autosomal chromosomes plus one pair of sex-determining chromosomes (XY in males and XX in females). The nuclear envelope begins to fragment and chromosomes are released in the cytoplasm. Meanwhile, the two centromeres separate from each other, migrating to opposite poles of the cell. The system of microtubules disassembles and reassembles, forming the mitotic spindle. In metaphase all chromosomes attach to the mitotic spindle via their kinetochores, forming an equatorial plate. Then, in anaphase, the centromeres of all chromosomes split, giving rise to haploid chromosomes, which migrate to the opposite poles of the mitotic spindle, taking a V-shape. In telophase two new nuclear envelopes grow from the ER, one around each set of chormosomes, and in citokinesis the membrane narrows in the equatorial plane, then the entire cell splits into two daughter cells.