Why can the mRNA strand made during transcription be thought of as a mirror image

If you're seeing this message, it means we're having trouble loading external resources on our website.

If you're behind a web filter, please make sure that the domains *.kastatic.org and *.kasandbox.org are unblocked.

Messenger RNA (mRNA) synthesis using DNA as the template is called transcription. The mRNA carries genetic information from the DNA of the chromosomes in the nucleus to the surface of the ribosomes in the cytosol. It is synthesized as a single strand. Chemically, RNA is similar to DNA. It is an unbranched linear polymer in which the monomeric subunits are the ribonucleoside 5′ monophosphates. The bases are the purines (adenine and guanine) and the pyrimidines (uracil and cytosine). Thymine is not used in mRNA. Instead, uracil is used. This base is not present in DNA. Messenger RNA is much smaller than DNA and is far less stable. It has a very short half-life (from seconds to minutes or hours) compared to that of nuclear DNA (years). Because it has a short half-life, the purine and pyrimidine bases that are used to make mRNA must be continually resynthesized. This requires the same array of nutrients noted previously for DNA synthesis.

The synthesis of mRNA from DNA occurs in several stages: initiation, elongation, editing (processing), and termination. Initiation of transcription (the synthesis of mRNA) occurs when factors that serve to stabilize nuclear DNA are perturbed. Perturbation signals pass in to the nucleus and stimulate transcription. A small portion of the DNA (∼ 17,000 bases) is exposed and used as the template for mRNA synthesis. The exposed portion also contains one or more sequences that have control properties with respect to the initiation of transcription. This region is called the promoter region and represents a key site for nutrient interaction. The promoter region precedes the start site of the structural gene and is said to be upstream of the structural gene. Those bases following the start site are downstream. The exposed DNA contains groups of bases called exons and introns. The introns are noncoding and are removed by editing prior to the movement of the mRNA from the nucleus to the cytosol.

Transcription is highly regulated. The DNA in all cell types is identical. However, not all of this DNA is transcribed in all cells all the time. Only certain genes are activated and transcribed into mRNA and subsequently translated into protein or peptides. As mentioned previously, these gene products give the individual cell type its identity. Central to this regulation are protein:DNA interactions and protein:nutrient interactions. At initiation, basal transcription factors recognize and bind to the start site of the structural gene. They form a complex with RNA polymerase II, an enzyme that catalyzes the formation of mRNA. Transcription factors bind to particular base sequences, called response elements, in the promoter region of the DNA that are upstream of the transcription start site (Figure 2). Each gene promoter contains a characteristic array of response elements, and these will determine to which signals the particular gene responds. Transcription factors also bind nutrients, and it is here that some nutrients have their effects on gene expression.

Why can the mRNA strand made during transcription be thought of as a mirror image

Figure 2. Schematic view of transcription.

The regulation of transcription often occurs through the regulation of transcription factors. These factors can be regulated by the rates of their synthesis or degradation, by phosphorylation or dephosphorylation, by ligand binding, by cleavage of a pro-transcription factor, or by release of an inhibitor. One class of transcription factors important for nutrition is the nuclear hormone receptor superfamily that is regulated by ligand binding. Ligands for these transcription factors include retinoic acid (the gene active form of vitamin A), fatty acids, vitamin D, thyroid hormone, and steroid hormones. These receptors are proteins with a series of domains. The retinoic acid receptor can serve as an example. Its ligand-binding domain recognizes and binds with high affinity the nutrient signal, retinoic acid. The DNA-binding domain gives gene specificity. It binds to a segment of the gene promoter that contains its corresponding response element, the retinoic acid response element (RARE). A transactivation domain then signals the effective occupation of this response element to the gene as a whole, including RNA polymerase II and its associated proteins. There are additional factors responsible for mediating this interaction between nutrient receptor and the transcription process. They include coactivating proteins, which stimulate transcription, and corepressor proteins, which can cause inhibition of transcription from a particular protein. In general, nutrients can signal the activation of transcription of some genes while at the same time turning off the transcription of others.

An interesting additional feature of this superfamily of nuclear hormone receptors is that they contain two zinc atoms in their DNA-binding domains. Each zinc is bound by four cysteine residues and causes the folding of the protein in a finger-like shape that binds DNA. The zinc ion plays an important role in gene expression because of its central use in the zinc finger of a wide variety of DNA binding proteins. In the case of the receptor superfamily, although zinc is required for receptor function, there is no evidence that it plays a regulatory role. However, there are other transcription factors in which it does play a role. MTF-1 (metal response element (MRE)-binding transcription factor-1) responds to increasing zinc concentrations within the cell by translocating to the nucleus and activating the transcription of genes containing MREs in their promoter region. These genes include metallothionein, which binds zinc and may play a key role in zinc homeostasis.

The direct binding of a nutrient signal to a transcription factor is perhaps one of the simpler ways in which nutrients impact gene transcription. There are other less direct but equally important mechanisms. Genes involved in cholesterol homeostasis are characterized by a sterol response element (SRE) in their promoter regions, which interacts with a sterol response element binding protein (SREBP). This protein is synthesized as a large precursor, incorporated into endoplasmic reticulum membranes, and is unavailable to function in gene regulation until it is cleaved and released. Limited cholesterol availability results in the cleavage and release of SREBP from the membrane compartment and its translocation to the nucleus. There it can perform its gene regulatory function by activating the transcription of genes for cholesterol synthesis as well as the LDL receptor gene. The LDL receptor facilitates cholesterol uptake by the liver. When it is abnormal due to a mutation in its gene, hypercholesterolemia results. The liver is unable to remove cholesterol from the blood and continues to synthesize it since SREBP remains active.

The metabolism and availability of macronutrients also influence gene transcription. Promoter elements have been described that allow a response to glucose (the carbohydrate response element (CHORE)). Although the specifics are unclear, the activity of the protein that binds this element responds to the metabolism of glucose and then stimulates the transcription of relevant genes—for example, those required for glucose metabolism (pyruvate kinase) and fatty acid synthesis (acetyl-coA carboxylase and fatty acid synthase). Fatty acids also influence gene expression. They can affect transcription by binding directly to their own transcription factor (the peroxisome proliferator activated receptor) and also indirectly by reducing the availability of SREBP within the nucleus. The latter mechanism provides a means for linking cholesterol and fatty acid metabolism with the cell.

Nutrients can also affect gene expression indirectly by regulating the release of hormones into the blood. Thus, glucose, in addition to having its own effects on gene expression through the CHORE, also stimulates insulin secretion from the pancreas. Insulin has its own transcriptional effects, often on the same genes that are regulated by glucose. In the postabsorptive state, insulin drops and glucagon is released. This hormone activates an intracellular signaling pathway that results in inhibition of genes involved in glucose metabolism and fatty acid synthesis and stimulation of genes involved in gluconeogenesis (e.g., phosphenolpyruvate carboxykinase). Taken as a whole, macronutrient availability regulates the expression of the complex set of genes responsible for macronutrient metabolism by an aggregate of direct and endocrine-mediated pathways.

View chapterPurchase book

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B0122266943022912

Neuroendocrinology: The Normal Neuroendocrine System

Zachary M. Weil, ... Donald W. Pfaff, in Progress in Brain Research, 2010

Chromatin: steroid hormone-induced covalent modifications of histone tails

New mRNA and protein synthesis in hypothalamic neurons are required for estrogenic facilitation of lordosis behavior (reviewed in Pfaff, 1980). This discovery led to a long series of experiments in which genes that had two properties were identified: (1) their mRNA levels were increased in hypothalamic neurons after E administration and (2) their gene products would foster female reproductive behaviors (summarized in Pfaff, 1999). In turn, the emphasis on gene–behavior relationships sprung two very surprising observations: the loss of one gene (ERalpha) could cause females to be treated like males and to respond like males (Ogawa et al., 1996), and secondly, an individual gene could have opposite behavioral effects in male brains as compared to female brains (summarized in Ogawa et al., 2004). All of these and other observations (e.g. Ogawa et al., 1998a, 1998b) caused us to focus on hormone-facilitated gene transcription.

The initiation of transcription is widely understood to involve changes in chromatin, the set of proteins that guard access to DNA by transcription factors such as ERs (illustrated in Figs. 3–5). Chromatin, the principal structural protein associated with DNA molecules, has a potential regulatory role in gene expression. Chromatin is the principal structural protein around which DNA is wrapped and it is made up of four canonical proteins termed H2A, H2B, H3 and H4, the structure of which have been largely conserved over evolutionary time. Broadly, DNA exists in two states, (1) euchromatin, wherein DNA is relatively loosely coiled around chromatin and which is associated with transcriptionally active genes. This is in contrast to (2) heterochromatin where DNA is tightly wrapped around histone peptides and thus the underlying DNA is inaccessible to the transcriptional machinery. These states were long considered developmentally programmed and static. However, these broad categories are significantly complicated by dozens of covalent modifications of histone tail regions which can more subtly regulate gene expression. Modification of histone tails with the additions of acetyl, phosphate, methyl, sumo and ADP ribosyl groups at specific amino acid residues alone and in combination has specific regulatory effects on the genes associated with modified chromatin. The complexity of this system has lead Professor C. David Allis and colleagues here at Rockefeller to suggest that these covalent modifications of histone peptides taken together form a ‘histone code’ which provide important epigenetic regulation of gene expression. Therefore, with respect to female sex behavior, it was natural for us to try and draw histone modification into E-caused transcriptional changes in the VMH neurons essential for E-dependent lordosis behavior (Weil et al., 2009). Results showed that treatment with 17β-estradiol rapidly methylates H3K9 in the VMH, and also that treatment with 17β-estradiol rapidly increased the acetylation status of histone H4 proteins in the VMH. These changes are highly likely to be relevant to lordosis behavior because double-identified ER-alpha positive cells in the ventromedial hypothalamus are shown by our new results to exhibit H4 acetylation. We have further reason to believe that histone acetylation within VMH neurons is causally linked to female reproductive behavior, because microinjections of the histone deacetylase inhibitor into the VMH potentiates estrogen-induced lordosis behavior and reduces rejection behaviors (Weil et al., 2009).

Why can the mRNA strand made during transcription be thought of as a mirror image

Fig. 3. Inactive DNA is tightly wound, whereas DNA accessible to transcription factors is no longer tied up in nucleosomes.

Why can the mRNA strand made during transcription be thought of as a mirror image

Fig. 4. Nucleosomes in which inactive DNA is wrapped are characterized by tails of histone proteins emerging in a manner that leaves them susceptible to chemical modification. In the usual case for mammalian cells, acetylation of amino acids on these tails is associated with increased rates of transcription and methylation is associated with repression of transcription. We have tied increased acetylation to estrogenic effects in the region of the ventromedial hypothalamus important for regulating lordosis behavior (Weil et al., 2009).

Why can the mRNA strand made during transcription be thought of as a mirror image

Fig. 5. The net result of loosening DNA tight connections with nucleosomes is thought to lie in freeing up transcription factor recognition elements, illustrated here by consensus nucleotide base sequences for estrogen receptors (ERs) and glucocorticoid receptors (GRs).

Similar considerations apply to maternal behavior. The optimal pattern of hormone exposure for the onset of maternal behaviors in female laboratory mammals comprises a sudden drop of progesterone levels in the blood coupled with high levels of estrogens in the blood. Surprisingly, however, it had not been determined exactly what duration of estrogen exposure would be minimally sufficient for the facilitation of maternal behaviors. We have just finished the appropriate experiments by ovariectomizing female mice 11 days before tests of maternal behaviors and, at the same time, implanting progesterone capsules subcutaneously, which will subsequently be removed two days before behavioral testing. The question is: how long must estrogens circulate in progesterone-withdrawn mice, for maternal behaviors to occur? The answer was surprising: estrogens administered as little as two hours before testing could enhance maternal behaviors in progesterone-withdrawn female mice (Murakami et al., unpublished data, 2010). The advantage of this short requirement of estrogen treatment is that it permits us to concentrate on a short time window during estrogen-triggered behaviorally important transcriptional changes. In turn, this knowledge tells us exactly when to look for histone chemical modifications in preoptic area neurons, a project that we are finishing now (Murakami et al., unpublished data, 2010). From unpublished experiments, we know that the schedule of E administration consistent with the initiation of maternal behaviors is associated with global changes of histone chemistry in the preoptic neurons essential for maternal behavior, and that these changes include both methylation and acetylation (Murakami et al., unpublished data, 2010), but we neither have tried to draw these histone modifications into transcriptional arguments, nor have we proven that these histone modifications are essential for the initiation of maternal behavior. Those experiments will be initiated soon in this laboratory.

Although the data and molecular concepts of histone modifications fit very clearly with access to estrogen-sensitive genes by ligand-activated transcription factors, ERs, phenomena in neuroendocrine neurons are not always so simple. The changes in methylation status of specific histones caused by acute stress in hippocampal neurons (Hunter et al., 2009) have not, so far, been mimicked by corticosterone injections as they might have been expected to be. Alternate routes of stress effects, therefore, are being explored currently.

Our current emphasis on the chemical modifications of histone tails in hypothalamic and preoptic neurons brings us through the next mechanistic step in demonstrating how gene–behavior causal relations occur. After all, Ogawa et al. proposed that deletion of a single gene could cause a female mouse to be treated like a male and to behave like a male (Ogawa et al., 1996). Altogether, knocking out the gene for ER-alpha caused a panoply of changes in sexual, aggressive and maternal behaviors (Ogawaet al., 1997, 1998a, 1998b). Taken together, the Ogawa et al. (2004) findings proved that the same gene could have opposite behavioral effects (on aggressive behavior) in females as that same gene has in males.

In sum, because of the clear relationships between patterns of gene expression and behavior in neuroendocrine systems, we predict that sex steroid effects on reproductive behaviors – both sexual and maternal – will provide some of the most chemically detailed and functionally important examples of histone modifications in nerve cells.

View chapterPurchase book

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/S0079612308810035

Regulation of Gene Expression

N.V. Bhagavan, Chung-Eun Ha, in Essentials of Medical Biochemistry (Second Edition), 2015

Regulation of mRNA Synthesis

In prokaryotes, mRNA synthesis can be controlled simply by regulating initiation of transcription. In eukaryotes, mRNA is formed from a primary transcript followed by a series of processing events (e.g., splicing, capping, polyadenylation). Eukaryotes regulate not only transcription initiation, but also the various later stages of processing.

An important aspect of mRNA regulation is determined by the degradation of mRNA molecules; i.e., translation can occur only as long as the mRNA remains intact. In bacteria, mRNA molecules have a lifetime of only a few minutes, and continued synthesis of mRNA molecules is needed to maintain synthesis of the proteins encoded in the mRNAs. In eukaryotes, the lifetime of mRNA is generally quite long (hours or days), thereby enabling a small number of transcriptional initiation events to produce proteins over a long period of time.

Metabolic pathways normally consist of a large number of enzymes; in some cases, the individual enzymes are used in a particular pathway and nowhere else. In these cases, it may be efficient to regulate expression of either all or none of the enzymes in the pathway. In bacterial systems, all enzymes of the pathway are encoded in a single polycistronic mRNA molecule called an operon, and synthesis of the single mRNA produces all the enzymes. In eukaryotes, common signals for transcription of different genes may be used, or the primary transcript can be differentially processed to yield a set of mRNA molecules, each of which encodes one protein. In eukaryotes, regulation of synthesis of the primary transcripts simultaneously regulates synthesis of all the gene products.

In prokaryotes, gene expression fluctuates in response to the environment because bacteria must be able to respond rapidly to a changing environment. However, due to the differentiation of cells of the higher eukaryotes, changes in gene expression are usually irreversible—for example, in the differentiation of a muscle cell from a precursor cell. Eukaryotes can change the number of copies of a gene during differentiation and, in this way, regulate the level of gene expression.

View chapterPurchase book

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780124166875000245

Nidovirales

In Virus Taxonomy, 2012

Synthesis of genomic and subgenomic rnas

Genome replication and sg mRNA synthesis (“transcription”) proceed through minus-strand intermediates. The genome serves as a template for the synthesis of full-length minus-strand RNA, from which in turn new genome copies are produced, but it is also believed to be the template for the synthesis of sg minus-strand RNA species (vide infra). The synthesis of viral RNAs is highly asymmetrical as plus-strand RNAs are produced in fast excess.

A hallmark of nidovirus transcription is the production of a 3′-coterminal nested set of sg mRNAs. The sg mRNAs of corona-, arteri- and bafiniviruses are chimeric, that is, comprised of sequences that are non-contiguous in the viral genome. Each carries a short 5′ leader sequence of 55–92, 170–210 nt, and 42 nucleotides, respectively, which is identical to the 5′ end of the viral genome. It was established early on that leader and “body” sequences are not joined through splicing, but via a process of discontinuous RNA synthesis. A key observation was the presence of mirror-copy nested sets of sg minus-strand RNAs in corona- and arterivirus-infected cells. Combined experimental evidence from biochemical and reverse genetics analyses indicates that these sg minus-strand RNAs are in fact the templates for sg mRNA synthesis. Replicative intermediates (RI)/replicative forms with sizes corresponding to the different sg mRNAs were shown to be actively involved in transcription. According to the prevailing 3′-discontinuous extension model, the discontinuous step occurs during the production of sg minus-strand RNAs and entails attenuation of RNA synthesis at the TRSs, followed by a similarity-assisted copy choice RNA recombination event. In corona-, arteri- and bafiniviruses, a TRS is present immediately downstream of the genomic leader sequence. It is believed that, during minus-strand RNA synthesis, the replicase complex upon encounter of an internal TRS dissociates from the template and is transferred to the 5′ end of the genome, guided by sequence complementarity between the anti-TRS on the nascent strand and the genomic TRS. Reinitiation and completion of RNA synthesis would then result in a chimeric minus-strand that in turn would serve as a template for uninterrupted (continuous) synthesis of 5′ leader-containing sg mRNAs.

Discontinuous sg RNA synthesis is not a trait of all nidoviruses. Ronivirus sg mRNAs lack a common 5′ leader and thus apparently arise from non-discontinuous RNA synthesis. Toroviruses employ a mixed transcription strategy; of the four sg RNAs, only RNA 2 carries a 15–18 nt 5′ leader derived from the 5′ end of the genome, whereas the others do not. It is likely that sg mRNAs are transcribed from sg minus-strand templates also in toro- and in roniviruses. Here, the conserved sequence elements (TPs) preceding the 3′-proximal genes might serve dual roles as signals for premature termination of minus-strand synthesis and as promoters for plus-strand production. The torovirus S gene, expressed from mRNA 2, lacks a TP. Apparently, transcription-competent minus-strand sg RNAs are produced by inclusion of a complementary copy of the 5′-terminal genomic TP via a similarity-assisted RNA recombination process analogous to that seen in corona- and arteriviruses.

Read full chapterView PDFDownload book

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780123846846000665

Deterministic Differential Equations

Paola Lecca, ... Thanh-Phuong Nguyen, in Computational Systems Biology, 2016

5.3.2 Transient Analysis

The transient analysis of deterministic differential models aims at determining the time-dependent values of the variables over a given time window.

Consider for instance the mRNA synthesis and degradation model given by Eq. (5.21). If we want to compute the value of the variable mRNA(t) for any t ≥ 0, we can directly solve the differential equation to find the following analytical closed-form expression for mRNA(t):

(5.26)mRNA(t)=k1k2+mRNA(0)−k1k2e−k2⋅t.

This function can be plotted for different values of the initial abundance of mRNA. Fig. 5.4 shows curves of the transient behavior of mRNA(t) for three different initial conditions when the rate constants k1 and k2 are set to 10.0 and 5.0, respectively. The plots in Fig. 5.4 allow us to see the convergence of the transient values to the steady-state solution that we determined in the previous section, and that is given by k1/k2 = 2.0. Notice that setting the initial abundance of the mRNA to the equilibrium value of 2.0 immediately nullifies the derivative in Eq. (5.26) and therefore provides a transient behavior that stays aligned with the steady-state value (curve for mRNA(0) = 0 in Fig. 5.4).

Why can the mRNA strand made during transcription be thought of as a mirror image

Fig. 5.4. Transient behavior of mRNA(t) over the time window [0,1] for the three different initial values mRNA(0) = 0, mRNA(0) = 2.0, and mRNA(0) = 3.0. The rate constants were set to k1 = 10.0 and k2 = 5.0.

The possibility of finding explicit solutions to systems of ordinary differential equations rapidly diminishes with the increase in model complexity. Therefore, the approach we described above, which allows us to determine the transient behavior of the modeled species by finding analytical closed forms for the continuous variables, cannot be generalized. Again, the transient analysis of models is usually done via numerical approximated methods, which starting from the initial known value of the variables (the value at time t = 0) compute an approximation for the values of the variables at time δt. This step-ahead computation is iteratively repeated to determine a polygonal curve approximating the real-valued variables over the time interval of interest in the analysis. Methods belonging to this family are usually (and improperly) referred to as numerical integration methods for ordinary differential equations.

A variety of approaches exist which differ on the amount of information used in computing the approximated value of the variable at time t + δt given its previously computed value at time t. The simplest one, which provides the essence of this numerical computation, is Euler’s method, which computes the value of the variable at time t + δt from the slope of the tangent line at time point t, by considering that the real unknown curve will not be significantly different from the tangent inside the interval (t − δt,t + δt) if a sufficiently small value of δt is used in the approximation step.

The approximation can be made arbitrarily precise by reducing δt, which, however, increases the number of evaluations required to determine the approximating polygonal in the interval of interest. A natural variation of Euler’s method is therefore one that considers an adaptive step, based on an estimation of the speed with which the distance of the tangent from the real value of the variable increases. Estimation of this distance requires the evaluation of higher-order derivatives for the variables. A commonly applied method of this type is the fourth-order Runge–Kutta method.

Numerical integration of systems of deterministic differential models provides a quick and effective means for transient analysis. Consider again the model described in Fig. 5.3 and the following differential equation model that describes the variation of the abundance of protein P over time:

(5.27)ddtP(t)=k1KnP(t)n+Kn−k2⋅P(t).

The first term on the right-hand side of Eq. (5.27) provides an abstract modeling of the repression effect exerted by the dimeric form of protein P on the synthesis of new molecules of protein P. Because the repression is realized by a cooperative effect (via the dimerized form of P), we can model the reduction of the synthesis rate by the negative Hill function, which provides a decreased synthesis rate as the abundance of P increases. Obviously, with such a complex form of the differential equation terms, it is not possible to find an explicit analytical solution for P(t). Resorting to numerical integration provides the time-dependent behavior shown in Fig. 5.5, where we plot the transient evolution of P(t) for three different values of the parameter K of the Hill function. Because K defines the amount of P that determines the half strength of the repression effect, increasing values of K result in less pronounced reductions in the synthesis rate, which results in increasing levels of protein P. In all cases, we assumed that the initial amount of protein P is zero — that is, P(0) = 0.

Why can the mRNA strand made during transcription be thought of as a mirror image

Fig. 5.5. Transient behavior of the P(t) over the time window [0,200] for three different values assigned to the Hill function parameter K: K = 0.3, K = 0.4, and K = 0.5. The other parameters in the differential equation model are set to the constant values k1 = 0.1, n = 2, and k2 = 0.01 and in all the three cases P(0) = 0.

Finally, we mention a special family of numerically integrators that have been developed to cope with stiffness, a common characteristic of differential equation models of biological systems. A model is stiff when it represents different phenomena which occur on very different timescales. In the model of a biological system, this is an easily encountered situation if we are representing events that proceed at very diverse speeds. For instance, consider the enzymatic reaction described in Eq. (5.13). The association and the dissociation of the enzyme and substrate is typically much faster than the catalysis step, and this is the main rationale that justifies our abstracting the two reactions to obtain the Michaelis–Menten form of the rate shown in Eq. (5.14). This means that inside the same model, some variables will be changing very fast and some will be changing much more slowly, which in turns calls for the necessity of our using very small integration steps to compute sufficiently accurate results. To solve stiff systems of ordinary differential equations, specific algorithms have been proposed, such as LSODA [212], which are implemented by most computational tools (see, eg, [208]) and are included in well-known numerical libraries such as ODEPACK [213].

View chapterPurchase book

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780081000953000056

Riboswitches as Targets and Tools

Pauline van Nies, ... Christophe Danelon, in Methods in Enzymology, 2015

2.5 Kinetics measurements by spectrofluorometry

Real-time monitoring of mRNA and protein synthesis was carried out on a fluorescence spectrophotometer (Cary Eclipse, Varian). The PUREfrex reaction solution was transferred to a 15-μL cuvette (Hellma) positioned in a Peltier thermostated four-cell holder maintained at 37 °C. Spinach and YFP fluorescence signals were detected every 30 s in the high-voltage mode at (excitation/emission wavelengths) 460/502 and 515/528 nm, respectively, to minimize fluorescence overlap.

For each conditions tested, at least three independent (no pooling of the PURE system reaction solution) fluorescence curves have been generated and representative kinetics are displayed (Fig. 4).

Why can the mRNA strand made during transcription be thought of as a mirror image

Figure 4. (A) Fluorescence excitation (dashed lines) and emission (solid lines) spectra of LL-Spinach (blue) and mYFP (green) measured in the PUREfrex expressing the mYFP-LL-Spinach gene. The LL-Spinach spectra were measured in the PUREfrex ΔR, that is devoid of ribosome, in the presence of 20 μM DFHBI. The mYFP spectra were collected in the PUREfrex without DFHBI. The arrowheads depict the excitation and emission wavelengths used for kinetics measurements. (B) Fluorescence intensity profiles of Spinach produced from the mYFP-LL-Spinach (dark blue) or YFP-Spinach (light blue) construct. (C) Apparent kinetics of mYFP (dark green) and YFP (light green) synthesis monitored simultaneously as in (B). (B and C) DNA concentration for both genes was 7.4 nM. (D) Plots of LL-Spinach fluorescence versus time using the mYFP-LL-Spinach (dark blue) or mYFP (light blue) construct. (E) Apparent kinetics of mYFP produced from the mYFP-LL-Spinach (dark green) or mYFP (light green) construct and monitored simultaneously as in (D). (D and E) DNA concentration for both genes was 0.74 nM. (F) Progression of Spinach fluorescence versus time in a PUREfrex ΔR (dark blue) or PUREfrex (light blue) reaction starting from 11.7 nM of the mYFP-LL-Spinach DNA. The arrowheads on the right axis point to the final intensity values used for calculating the conversion factor between fluorescence a.u. and mRNA concentration. (G) Apparent kinetics of mYFP synthesis in a PUREfrex ΔR (dark green) or PUREfrex (light green) reaction monitored simultaneously as in (F). (F and G) The arrow at around 100 min indicates the addition of DNaseI to stop transcription. (D and F) Concentrations of mRNA were calculated using a conversion factor of 10 nM/a.u. (E and G) Concentrations of mYFP were calculated using a conversion factor of 0.33 nM/a.u.

The cuvettes need to be thoroughly cleaned immediately after each experiment to maximize data reproducibility. Washing was done by successively filling the cuvette with Hellmanex 2%, KOH 1 M, nuclease-free water, EtOH 100% for 1 min in a bath sonicator, applying three washes with nuclease-free water in between the different reagent treatments.

View chapterPurchase book

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/S0076687914000494

Nuclear Architecture and Transcriptional Regulation of MicroRNAs

Pavithra L. Chavali, Sreenivas Chavali, in MicroRNA in Regenerative Medicine, 2015

43.5 DNA-Dependent RNA Transcription

From the previous sections, it becomes clear that the chromatin structure imposes profound and ubiquitous effects on almost all DNA-related metabolic processes, including transcription, recombination, and DNA repair and replication. In this section, the main focus is on the steps in transcription regulation, with emphasis on chromatin structure. The regulatory steps are similar to both mRNA and microRNAs (miRNAs) until 3′ end processing, since most of the protein-coding genes and miRNAs undergo RNA Pol II–mediated transcription.

The typical Pol II transcription cycle for a naked DNA molecule begins with the binding of activators upstream of the core promoter, including the TATA box and transcription start site (Table 43.1). This event leads to recruitment of adaptor complexes, such as SAGA, with multiple enzymatic activities [25], which leads to the binding of the TATA binding protein (TBP) to the TATA box, facilitating the assembly of TFIID upstream of the core promoter. After TFIID binds to the TATA box via the TBP, five more TFs and RNA polymerase combine around the TATA box in stages to form a closed preinitiation complex (PIC). TFIIH then melts 11–15 bp of DNA to position the single-strand template in the Pol II cleft (open complex), which initiates RNA synthesis. The carboxy-terminal domain (CTD) of Pol II is phosphorylated by the TFIIH subunit during the first 30 bp of transcription and loses its contacts with the general TFs before the elongation stage.

Table 43.1. Characteristics of Common Promoter Elements

Promoter ElementDistance from TSSConsensusFunctionTATA box28–34 bp upstreamTATAABinding of PIC complexInitiator (Inr)at +1 TSS siteYYANWYYRecruitment of PIC complexDownstream promoter element (DPE)28–32 bp upstreamRGWYYDirect PIC to TSSTFIIB recognition element (BRE)Upstream of TATA boxSSRCGCCIncrease or decrease in transcriptionCpG islandTSS, intragenic, or intronic(CG)nTranscription of ubiquitous genes

However, the PIC alone drives only the basal rate of transcription. Other proteins known as activators and repressors, along with any associated co-activators or co-repressors, are responsible for modulating transcription rate. The phosphorylated CTD begins to recruit the factors that are important for productive elongation and mRNA processing [26]. While TFs can share targets and bind to sequence-specific binding sites in the context of free DNA, they have to adopt a different strategy to achieve binding to sites buried in chromatin. Early biochemical experiments suggested that TFs can bind to nucleosomal DNA in a cooperative manner before nucleosome disassembly. However, later studies showed that chromatin-remodeling complexes help in recruiting TFs to chromatin sites [27]. Additionally, it has been demonstrated that promoters have a lower density of nucleosomes than the coding regions [28].

Considering the amount of DNA directly contacted by Pol II and general TFs, the structure of the nucleosome seems to pose a significant obstacle in PIC formation [29]. A number of studies have proposed that, upon gene activation, nucleosomes are dissembled and reassembled as the transcription is completed. In other cases, nucleosomes are not displaced; rather, promoters assemble partial PICs, excluding Pol II and TFIIH. This implies that when template DNA engages into the Pol II active site, DNA–histone contacts may be broken (Figure 43.4(a,b)).

Why can the mRNA strand made during transcription be thought of as a mirror image

FIGURE 43.4. Life cycle of messenger RNA.

(a) Chromatin in a condensed state is remodeled by chromatin remodelers with the aid of histone-modifying enzymes like histone acetyl transferases and histone deacetylases, which induce an open chromatin conformation.

(b) The open chromatin conformation invokes the assembly of a preinitiation complex (PIC) consisting of Pol II on the chromatin’s open region.

(c) Pol II and associated TFs, including TFIIA, TFIIB, TFIID, TFIIE, TFIIF, and TFIIH, assemble on the promoter to form the PIC (left). Mediator, a large multi-subunit complex, regulates PIC assembly. Next TFIIH, a complex harboring DNA helicases, melts the DNA to expose the template strand, marking the beginning of RNA synthesis. Once the first nucleotide bonds have been formed, Pol II is released from the promoter to facilitate downstream transcription (right). Most of the general TFs dissociate from the promoter, whereas mediator is likely to remain associated at the promoter to facilitate the next round of polymerase recruitment and preinitiation. The first modification to the precursor mRNA is an m7GpppN cap structure added at the 5′ end of the nascent transcript. In the nucleus, this structure recognizes the cap-binding complex (CBC), consisting of CBP20 and CBP80. The CBC-bound cap structure protects the nascent transcript from nuclease attack and plays important roles in the export of mature mRNAs from the nucleus.

(d) Most eukaryotic genes contain introns, which because they are non-protein-coding, are removed from the template by splicing. After splicing, the EJC is deposited ≈24 nt upstream of the exon–exon junctions. EJC is important in mRNA export from the nucleus and in cytoplasmic mRNA quality control.

(e) Added to the 3′ end of the transcript is a poly(A) tail, approximately 200−300 nt long, that facilitates binding of the poly A binding protein (PABP). The tail aids transcription termination and mRNA export from the nucleus, and protects the mRNA from nuclease activity in the cytoplasm and in translation initiation.

(f) Tap/p15, along with SR proteins (splicing proteins rich in serine and arginine), exports mRNA to the nuclear pore complex (NPC).

(g) On the cytoplasmic side of the NPC, complexes containing Dbp5 remove RNA binding proteins from the mRNA, ensuring that its export to the cytoplasm is one way.

(h) In the cytoplasm, the cap structure is the binding site for eIF4E recruitment of the ribosome. The poly(A) bound PABP interacts with 5’-cap-bound eIF4G and eIF4E to synergistically bring about translation initiation and facilitate ribosome binding. A sequence of events then leads to polypeptide.

(i) Deadenylases shorten the mRNA poly(A) tail, aiding exosomes in mRNA degradation.

(j) Protein synthesis is completed by post-translational modification(s), proper 3D polypeptide folding, and subsequent cellular localization of resultant proteins to their functional sites.

Transcription elongation begins when Pol II releases from general TFs and travels into the coding region. This event signals the recruitment of the elongation machinery, which includes the factors involved in polymerization, mRNA processing, mRNA export, and chromatin function. The main difference between the factors that aid initiation and those that aid elongation is that the elongation factors require direct or indirect binding to the CTD of Pol II [26]. The Pol II CTD undergoes two major phosphorylation changes during elongation: Serine5 is phosphorylated by TFIIH at the 5′ end of the open reading frame (ORF), and Serine2 is phosphorylated by the CTK kinase, as Pol II transits toward the 3′ end. This controls and couples elongation with alterations in chromatin structure. For example, an evolutionarily conserved multi-subunit complex called PAF is loaded onto the Serine5-phosphorylated CTD and regulates the binding of other CTD-associated chromatin regulators, such as the histone H3K4 methyltransferase Set1 complex for elongating Pol II [30]. In contrast, the histone methyltransferase Set2 targets primarily Serine2-phosphorylated CTD while Pol II travels toward the 3′ end of the ORF [31].

The first change that a nascent pre-mRNA transcript undergoes is 5′ capping. When a transcript reaches about 20–30 nt in length, a 7-methylguanosine cap is added to the 5′ end, which protects the nascent pre-mRNA from degradation by the cap-binding complex (CBC) [32] (Figure 43.4(c)). Pre-mRNA capping is followed by splicing. Although the exons are relatively small and embedded in large intron sequences, the splicing machinery recognizes the exons, removes the introns from the pre-mRNA molecule, and ligates the exons to form a mature mRNA. The main features of intron in splicing include (1) the exon–intron junction at the 5′ and 3′ ends of introns (5′ ss (single strand ) and 3 ss); (2) the branch site sequence located upstream of the 3′ ss; and (3) the polypyrimidine tract located between the 3′ ss and the branch site. All types of pre-mRNA splicing take place within the spliceosome—a large complex composed of five small nuclear RNA (snRNA) molecules (U1, U2, U4, U5, and U6)—and as many as 150 proteins. Each of the five snRNAs assembles with proteins to form small nuclear ribonuclear protein complexes (snRNPs) [33].

Once splicing is completed, the exon-junction complex (EJC) proteins are deposited at the site of exon fusion (Figure 43.4(d)). Capping and splicing are both important for the recruitment of the transcription-export (TREX) complex. This is primed by the 3′ end processing of transcripts, aided by transcription termination. This step is crucial not only to release Pol II from its template but also to ensure that a pool of polymerases is available for reinitiation or new transcription. Termination can also prevent formation of antisense RNAs, which can interfere with normal pre-mRNA production, thereby preventing aberrant gene expression. Transcription termination can occur anywhere from a few base pairs to several kb downstream from the 3′ end of the mature mRNA. This process is coupled to 3′-end processing of the pre-mRNA, namely polyadenylation, which refers to the addition of a poly(A) tail.

In mammals, cleavage and polyadenylation occur 10–30 nts downstream from a conserved hexanucleotide, AAUAAA, and 30 nt upstream of a less conserved U- or GU-rich region. This signal is recognized by the cleavage and polyadenylation specificity factor (CPSF). CPSF binds other pre-mRNA binding protein components such as poly(A) polymerase and poly-A binding protein (PABP) (Figure 43.4(e)). PABP protects the pre-mRNA from exonuclease degradation by binding the poly(A) tail, and it is required for correct and efficient poly(A) tail synthesis [34]. Several other factors, including the CTD of polymerase help link 3′-end processing to transcription. Another interesting observation is that H3K36 methylation often decreases near the poly(A) site during or prior to Pol II release, suggesting that H3K36 methylation plays a role in transcription termination [35].

The final step in mature mRNA synthesis is the nuclear export of mRNA, now bound with protein components and called mRNP. Following completion of proper nuclear processing and the recruitment of an export receptor, an mRNP is considered to be export-competent. The TREX complex is highly conserved and is essential for mRNP export (Figure 43.4(f)). This export-competent mRNP is targeted to the nuclear pore complex (NPC) via its export receptor. For some transcribed genes, the positioning of the respective chromatin region near the NPC might facilitate export by physically linking the processes. The export receptor docks at the NPC via special proteins called FG-Nups that mediate the movement of the mRNP through the NPC by some type of facilitated diffusion. When the mRNP reaches the cytoplasmic site of NPC, specific proteins from mRNPs are lost by ATP hydrolysis and recycled for the next round of mRNA binding. The binding of cytoplasmic factors occurs immediately with the entry of the 5′ end of the transcript into the cytoplasm; the CBC is replaced with eIF4e, ribosomes bind, and translation begins before the entire mRNP has been extruded from the NPC (Figure 43.4(g–j)).

Section Key Points

RNA Polymerase II transcribes mRNA, which undergoes nuclear processing, including 5′ capping, splicing, and 3′ polyadenylation before being exported into the cytoplasm to be translated.

View chapterPurchase book

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780124055445000435

Antituberculosis drugs

Peter R Donald, Helen McIlleron, in Tuberculosis, 2009

RIFAMPICIN

Mode of action

Rifampicin prevents DNA-directed mRNA synthesis by binding to RNA polymerase. The mutations responsible for the majority of rifampicin resistance lie in the rpoB gene and occur with a frequency of 1 in 107 to 1 in 108 cell divisions.22 The MIC of rifampicin is 0.25 μg/mL in broth and 0.5 μg/mL on agar.

Pharmacokinetics

Rifampicin reaches peak concentrations 1.5–4 hours after oral administration. Following a meal, absorption is delayed and reduced by about 20%.23,24 It is highly (80–90%) protein bound but, nevertheless, penetration into lung tissue, tuberculous cavities and kidneys is good, reaching concentrations higher than that in serum. Less satisfactory concentrations are achieved in pyogenic bone lesions, pleural empyema and cerebrospinal fluid.25

Following a dose of 600 mg, peak concentrations of 6–14 μg/mL can be expected at the start of treatment.5 Rifampicin is converted principally to the active 25-desacetyl metabolite. Metabolism is self-induced, resulting in reductions in the area under the concentration–time curve of 25–45% after repeated doses.5,26 Elimination is primarily hepatic and competition with bilirubin for hepatic excretion may cause transient hyperbilirubinaemia. A serum concentration of rifampicin between 8 and 24 μg/mL 2 hours after drug administration has been recommended,14 but a number of studies in patients have shown widely variable serum concentrations with median 2-hour concentrations of less than 8 μg/mL. HIV infection, diabetes, male sex, alcohol use and undernutrition are factors reported to predispose patients to lower rifampicin concentrations.15,27–29 Significant differences in bioavailability between various rifampicin-containing formulations on the market have also been reported.30

Clinical efficacy

At a 600-mg dose, rifampicin has moderate bactericidal activity, approximately half that of isoniazid,5 but its unique ability is to sterilize TB lesions within 6–9 months. This is thought to be due to the particularly rapid onset of action of rifampicin which enables it to kill intermittently metabolically active bacilli.31 Acting with pyrazinamide, rifampicin is essential for the efficacy of modern 6-month, short-course therapy.6 Rifampicin should be administered daily during the intensive phase of treatment as intermittent dosing is associated with delayed culture conversion and with acquired rifamycin resistance in patients coinfected with HIV.32

The loss of rifampicin as an effective agent when resistance to it develops constitutes a major setback for patients and TB control programmes.

Studies of the EBA of rifampicin at a dose of 600 mg show that it is moderately active with a 2-day reduction of bacillary numbers of approximately 2.0 log10 cfu/mL sputum/day. When given in a two-drug combination with isoniazid it allowed isoniazid resistance to emerge in 0.5% of patients.21

Side effects

Rifampicin potently induces the expression of several proteins that affect the metabolism of other drugs. These proteins include microsomal enzymes (e.g. cytochrome P450 3A4/5, 2A6, 2C8/9/19, 2B6), phase II enzymes (e.g. UDP-glucuronosyltransferases and glutathione-S transferases) and drug transporters (e.g. p-glycoprotein) amongst others. As a result, the concentrations of some drugs are profoundly reduced. Thus administration of rifampicin together with other drugs may result in reduced efficacy of the concomitant drug, or increased toxicity if the concentrations of a toxic metabolite are increased (Appendix 2).

The hepatotoxicity of isoniazid, pyrazinamide and a variety of other drugs, including non-anti-TB agents, may be potentiated by rifampicin, and rifampicin-associated hepatoxicity is more likely in patients with chronic liver disease. Other severe reactions to rifampicin are rare; these are usually associated with sensitization and are more common with intermittent doses. Patients should be warned that the orange discoloration of body fluids caused by the rifamycins can permanently stain clothing and soft contact lenses. The occurrence of thrombocytopenia is considered to be an absolute contraindication to using rifampicin again. Side effects and drug–drug interactions are listed in Appendix 2.

View chapterPurchase book

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9781416039884000597

PHOTOSYNTHESIS AND PARTITIONING | C3 Plants

A.S. Raghavendra, in Encyclopedia of Applied Plant Sciences, 2003

Regulation of Gene Expression

There is a perfect coordination of mRNA synthesis and protein accumulation during leaf development and chloroplast maturation. Much of the control has been found at the level of gene transcription. Light has a central role in regulating the biogenesis of the photosynthetic apparatus during chloroplast development, including switching on the expression of genes encoding enzymes of C3 photosynthesis. These processes appear to be complex, involving many regulatory proteins and cross-talk between the different photoreceptor signaling pathways.

The expression of C3 photosynthesis genes in leaves is also modulated by environmental and metabolic signals. High levels of glucose or sucrose reduce levels of a number of Calvin cycle mRNAs, including those encoding the small subunit of rubisco, SBPase, and FBPase. This feedback mechanism, acting at the level of transcription, might be important for source–sink regulation in the plant. Photosynthesis genes respond remarkably to nutrient status, particularly nitrogen and phosphorus levels, and this is influenced again by the carbohydrate status. Thus, an interaction between carbohydrate and nutrient status may be a long-term strategy to control carbon metabolism.

Why can the mRNA strand made during transcription be thought of as a mirror image of the DNA strand from which is was made?

Why can the mRNA strand made during transcription be thought of as a mirror image of the DNA strand from which it was made? the mRNA strand made during transcription can be thought of as a mirror image of the DNA strand from which it was made is complementary to the DNA molecule but it is not identical to it.

What happens when mRNA is transcribing DNA?

During transcription, the enzyme RNA polymerase (green) uses DNA as a template to produce a pre-mRNA transcript (pink). The pre-mRNA is processed to form a mature mRNA molecule that can be translated to build the protein molecule (polypeptide) encoded by the original gene.

Why does mRNA have to be copied in transcription?

A copy of mRNA must be made during transcription to effect translation of protein molecule with specific sequence of amino acids on ribosomes as per the codons in DNA molecule.

What direction is the mRNA transcript manufactured?

RNA polymerase synthesizes an RNA transcript complementary to the DNA template strand in the 5' to 3' direction. It moves forward along the template strand in the 3' to 5' direction, opening the DNA double helix as it goes.