6.4 CHEMICAL AND THERMODYNAMIC PROPERTIES OF NUCLEIC ACIDS

To understand how nucleic acids function, we must understand their chemical properties as well as their structures. The role of DNA as a repository of genetic information depends in part on its inherent stability. The chemical transformations that do happen are generally very slow in the absence of an enzyme catalyst. The long-term storage of information without alteration is so important to a cell, however, that even very slow changes in DNA structure can be physiologically significant. Other, nondestructive alterations of DNA do occur and are essential to function, such as the strand separation that must precede replication or transcription. For RNAs, chemical modifications can play significant roles in ensuring correct structure and function.

196

In addition to providing insights about physiological processes, our understanding of nucleic acid chemistry gives us a powerful array of technologies that have applications in molecular biology, medicine, and forensic science. We examine here the chemical properties of DNA and RNA, as well as some of these technologies. Many more techniques and applications are discussed in Chapter 7.

Double-Helical DNA and RNA Can Be Denatured

Solutions of carefully isolated, double-stranded DNA are highly viscous at pH 7.0 and room temperature (25°C). When such a solution is subjected to extremes of pH or to temperatures above 80°C, its viscosity decreases sharply, indicating that the DNA has undergone a physical change. This change is due to denaturation, or melting, of the double-helical DNA and can also occur with RNA. Disruption of both the hydrogen bonding between paired bases and the base stacking causes the double helix to unwind, forming two single strands that are completely separate from each other along the entire (or partial) length of the molecule. No covalent bonds in the nucleic acid are broken during denaturation (Figure 6-28).

Figure 6-28: Reversible denaturation of DNA. DNA is shown here, but RNA is also capable of denaturation and reannealing.

Renaturation of a DNA or RNA molecule is a rapid, one-step process, as long as a double-helical segment of at least a dozen residues still unites the two strands. When the temperature or pH is returned to the range in which most organisms live, the unwound segments of the two strands spontaneously rewind to yield the intact duplex. This process, called annealing, involves re-formation of all the base pairs in the double helix. If the strands were completely separated, renaturation occurs in two steps. In the first step, which is relatively slow, complementary sequences in the two strands “find” each other by random collisions and form a short segment of double helix. The second step is much faster: the remaining unpaired bases successively come into register as base pairs, and the two strands “zipper” themselves together to form the double helix.

197

The close interaction of stacked bases in a nucleic acid has the effect of decreasing its absorption of UV light relative to that of a solution with the same concentration of free nucleotides, and the absorption is further decreased by the pairing of two complementary strands. Hydrogen bonding and base stacking in the double helix limit the resonance of the aromatic rings of the bases, thereby decreasing UV light absorption. This is known as the hypochromic effect (Figure 6-29). When DNA is denatured, the base pairs are disrupted and the two strands separate into randomly coiled chains. The resonance of the bases in each strand is no longer constrained, as it is when the bases are part of a double helix. As a result, the UV light absorption of single-stranded DNA is approximately 40% higher than that of double-stranded DNA at the same concentration. This increase in absorption is called the hyperchromic effect. The transition from double-stranded DNA to the single-stranded, denatured form can thus be detected by monitoring the absorption of UV light.

Figure 6-29: UV light absorption by DNA. The transition from double-stranded DNA to the single-stranded, denatured form can be detected by monitoring UV light absorption by the sample (shown here as A260, absorbance at a wavelength of 260 nm). The melting point (Tm) is the temperature at which half the DNA in the sample is denatured.

DNA molecules in solution denature when they are heated slowly. Each species of DNA has a characteristic denaturation temperature, or melting point (Tm), defined as the temperature at which half the DNA is denatured. In general, the higher the content of G≡C base pairs in the DNA, the higher its melting point. This is because G≡C base pairs, with three hydrogen bonds, require more heat energy to dissociate than do A=T base pairs. Careful determination of the melting point of a DNA specimen, under fixed conditions of pH and ionic strength, can yield an estimate of its base composition. If denaturation conditions are carefully controlled, regions that are rich in A=T base pairs will specifically denature while most of the DNA remains double-stranded. Such denatured regions, or bubbles, can be visualized with electron microscopy (Figure 6-30). Strand separation of DNA must occur in vivo during processes such as DNA replication and transcription. As we will see in Chapter 11, the DNA sites where these processes are initiated are often rich in A=T base pairs.

Figure 6-30: Partially denatured DNA. The DNA shown in this electron micrograph was partially denatured, then fixed to block renaturation during sample preparation. The arrows point to some single-stranded bubbles where denaturation has occurred. The regions that denature are reproducible in repeated experiments and are rich in A = T base pairs.

RNA duplexes or RNA-DNA hybrid duplexes can also be denatured. Notably, RNA duplexes are more stable than DNA duplexes. At neutral pH, the denaturation of a double-helical RNA often requires higher temperatures, by 20°C or more, than those required to denature a DNA molecule with a comparable sequence. The stability of an RNA-DNA hybrid is generally intermediate between that of double-stranded RNA and DNA.

198

Nucleic Acids from Different Species Can Form Hybrids

The ability of two complementary DNA strands to pair with each other can be used to detect similar DNA sequences in different species or within the genome of a single species. If double-stranded DNAs isolated from human cells and from mouse cells are completely denatured by heating, then mixed and kept at 25°C for many hours, much of the DNA will anneal. Most of the mouse DNA strands anneal with complementary mouse DNA strands to form mouse duplex DNA; similarly, most human DNA strands anneal with complementary human DNA strands. However, some strands of the mouse DNA will associate with human DNA to yield hybrid duplexes, in which segments of a mouse DNA strand form base-paired regions with segments of a human DNA strand (Figure 6-31). The ability to hybridize DNA from different species is a valuable laboratory tool for exploring evolutionary relationships. Different species have proteins and RNAs with similar functions—and often similar structures. In many cases, the DNAs encoding these proteins and RNAs have similar sequences. The closer the evolutionary relationship between two species, the more extensively their DNAs will hybridize. For example, human DNA hybridizes much more extensively with mouse DNA than with DNA from yeast.

Figure 6-31: Cross-species DNA hybridization. Two DNA samples can be compared by heating a mixture of the DNAs to denature the strands and then cooling the mixture to allow complementary strands to form duplexes. The greater the sequence similarity between the two DNA samples, the more hybrid duplexes will form, with one strand derived from the first species and the other from the second.

The hybridization of DNA strands from different sources forms the basis for powerful techniques used in classical molecular genetics. Although these hybridization techniques are used less often as new, high-throughput approaches based on DNA sequencing become available (as described in Chapters 7 and 8), they are still widely used in research laboratories worldwide. Such techniques have laid the foundation for understanding the study of nucleic acids.

A specific gene’s DNA or RNA sequence can be detected in the presence of many other sequences by hybridization with a probe, a carefully chosen nucleic acid sequence complementary to the gene of interest. To be visualized in the laboratory, the probe must be labeled in some way, usually radioactively or with a fluorophore (a compound carrying a fluorescent group). The probe that is selected depends on what is known about the gene under investigation. Sometimes, a gene from another species that has sequence similarity to the gene of interest makes a suitable probe. If the protein product of a gene has been purified, probes can be designed and synthesized by working backward from the amino acid sequence, deducing the DNA sequence that would code for it. Or, researchers can often obtain the DNA sequence information necessary for creating a probe from sequence databases that detail the structure of millions of genes from a wide range of organisms. Because base-pairing stability is sensitive to pH and temperature, these parameters can be adjusted experimentally to detect nucleic acid sequences with varying degrees of complementarity to the probe. The technique is sensitive enough to reveal sequences that differ by a single base pair. This can be critically important in medical and forensic applications, given that two people can share genetic information that differs at only one or a few base pairs, called single-nucleotide polymorphisms (SNPs).

199

In other common hybridization methods, gel electrophoresis is used to separate DNA or RNA molecules by size (Figure 6-32a). A variation of gel electrophoresis, used to detect proteins under denaturing conditions, was discussed in Chapter 4 (see Highlight 4-1). Here, the gel matrix is not denaturing but instead is made of agarose, a kelp-derived material that does not disrupt nucleic acid base pairing. The starting nucleic acid sample, in solution in a test tube, is applied to a slot at one end of the gel, and a voltage is applied. Because DNA and RNA molecules are negatively charged, they migrate toward the positive end of the gel matrix in the electric field. Larger molecules tend to move more slowly than smaller ones, so this provides a means of separating nucleic acids by size. Following electrophoresis, the DNA fragments are transferred to a nitrocellulose membrane so that their positions in the gel relative to each other are preserved on the membrane. Once on the membrane, the nucleic acid can be hybridized with a DNA or RNA probe, labeled so that it can be detected by measuring radioactivity or fluorescence.

Figure 6-32: Gel electrophoresis used in the Southern and Northern blotting techniques. (a) Gel electrophoresis is used to size-fractionate a DNA (Southern blotting) or RNA (Northern blotting) mixture. The samples are then transferred to (i.e., are blotted onto) a nitrocellulose membrane, where they are detected using short radiolabeled oligonucleotide probes that base-pair to the samples on the membrane. (b) Northern blot analysis of RNA isolated from various human tissues. For each sample, approximately 10 μg of total RNA was separated on a 1.2% agarose-formaldehyde gel, transferred to a membrane, and hybridized to a 32P-labeled probe—an mRNA for human platelet endothelial cell adhesion molecule (PECAM-1). The same blot was also probed with a cDNA (complementary DNA, a DNA copy of an mRNA sequence; see Chapter 7) for glyceraldehyde 3-phosphate dehydrogenase (GAPDH) to control for the amount of material in each lane. (GAPDH mRNA is used as a control because it is found in all tissues, in almost equal amounts.) Note the differences in PECAM-1 RNA levels detected in the different tissues; two bands are observed for PECAM-1 in each lane because there are two distinct forms of the mRNA for this gene.

When used to detect DNA, this method is known as Southern blotting, named for Edwin Southern, who invented the technique at the University of Edinburgh. When used for detecting RNA, the technique is called Northern blotting, because of its similarity to the Southern method. Applications of these techniques include identifying a person on the basis of a single hair left at the scene of a crime or predicting the onset of a disease in an individual decades before symptoms appear. Northern blotting can also be used to detect the levels of a particular type of RNA in different body tissues (Figure 6-32b), providing fascinating insight into how cells regulate the expression of genes. Notably, these classical methods of Southern and Northern blotting are still used to answer specific experimental questions, despite the development of high-throughput strategies based on DNA sequencing technology (see Chapters 7 and 8).

Nucleotides and Nucleic Acids Undergo Uncatalyzed Chemical Transformations

Purines and pyrimidines, and the nucleotides of which they are a part, can undergo spontaneous alterations in their covalent structure. The rate of these reactions is generally very slow, but as noted earlier, they are physiologically significant because of the cell’s low tolerance for changes in its genetic information. Alterations in DNA structure that produce permanent genetic changes are known as mutations. Extensive evidence suggests an intimate link between the accumulation of mutations in an individual organism and the processes of aging and carcinogenesis.

200

Several nucleotide bases undergo deamination, a spontaneous loss of their exocyclic amino groups. For example, under typical cellular conditions, deamination of cytosine (in DNA) to uracil occurs in about 1 of every 107 C residues in 24 hours (Figure 6-33). This corresponds to about 100 spontaneous deamination events per day, on average, in a mammalian cell. Deamination of adenine and guanine occurs at about 1/100th this rate.

Figure 6-33: Cytosine deamination to uracil. Only the base is shown.

The slow cytosine deamination reaction seems innocuous enough, but it is almost certainly the reason that DNA contains thymine rather than uracil. In DNA, uracil is the product of cytosine deamination, and it is readily recognized as foreign and is removed by a DNA repair system (see Chapter 12). If DNA normally contained uracil, the recognition of U residues resulting from cytosine deamination would be more difficult, and unrepaired uracils would lead to permanent sequence changes as they were paired with adenines during replication (cytosine normally pairs with guanine, so introduction of uracil into DNA effectively changes a C≡G base pair to a U–A base pair). Establishing thymine as one of the four bases in DNA may well have been a crucial turning point in evolution, making the long-term storage of genetic information possible.

Cytosine deamination also provides innate cellular defense against viral infection. A family of human proteins called APOBECs catalyze cytosine deamination in the viral genome during the initial round of replication by HIV. This hypermutation results in many nonviable viral particles, eventually destroying the coding capacity of the virus. In HIV and related viruses, the viral protein Vif binds to APOBECs and triggers their degradation. Vif has therefore become an important antiviral target, because viruses lacking this protein are much less capable of establishing chronic infection in human cells.

Base Methylation in DNA Plays an Important Role in Regulating Gene Expression

Certain nucleotide bases in DNA molecules are enzymatically methylated, usually after DNA synthesis is complete. Adenine and cytosine are methylated more frequently than guanine (Figure 6-34a). Methylation is generally confined to certain sequences or regions of a DNA molecule. For example, more than half of all CpG sequences in mammalian genomes are methylated on the C residue. Methylation tends to inhibit gene expression, because the methylated DNA is not efficiently copied into RNA. In many cancers, gene regulatory regions in DNA become abnormally hypermethylated. This can result in the silencing of genes that would otherwise control cell growth. DNA methylation may affect gene transcription by physically blocking the binding of proteins that facilitate transcription. Other proteins, however, can specifically bind to methylated DNA and recruit additional proteins that help form compact, inactive regions of chromosomal DNA.

Figure 6-34: Chemical modifications in DNA and RNA. (a) Modified nucleotides in DNA. The most common postsynthetic modification to DNA is base methylation. 5-Methyldeoxycytidine occurs in the DNA of animals and higher plants; the other methylated bases shown here are produced by specific enzymes. (b) Modified nucleotides in RNA. Enzyme-catalyzed RNA base modifications are common in tRNA and rRNA, although the function of such alterations is not always clear. The presence of N4-acetylcytidine in bacterial tRNAs may enhance protein synthesis.

201

All known DNA methylases (methyltransferases) use S-adenosylmethionine as a methyl group donor (see Figure 6-13). E. coli has two prominent methylation systems. One serves as part of a defense mechanism that helps the cell distinguish its DNA from foreign DNA by marking its own DNA with methyl groups and destroys foreign, nonmethylated DNA, a process known as restriction modification (discussed in Chapter 7). The other system methylates A residues in the sequence 5′-GATC-3′ to form N6-methyldeoxyadenosine. Methylation in this case is mediated by the Dam (DNA adenine methylation) methylase, a component of a system that repairs the mismatched base pairs that occasionally form during DNA replication (see Chapter 12).

In eukaryotic cells, about 5% of all C residues in DNA are methylated to form 5-methyldeoxycytidine (see Figure 6-34a). As noted above, methylation is most common at CpG sequences, producing methyl-CpG symmetrically on both strands of the DNA. The observed 5% methylation frequency of C residues is lower than would be expected based on the random presence of CpG sequences in a genome, and, in fact, CpG dinucleotides occur much less often in vertebrate genomes than predicted by chance. For example, in the human genome, which has a G≡C content of ∼40%, the frequency of CpG would be predicted as 0.2 × 0.2 = 0.04, or 4%, but the frequency is closer to just 1%. The extent of methylation of CpG sequences varies with molecular region in large eukaryotic DNA molecules. In addition to the tendency of methylation of C residues to inhibit gene expression, methylation suppresses the migration of segments of DNA called transposons (see Chapter 14).

KEY CONVENTION

When a chemical group attached to an atom in the purine or pyrimidine ring is altered, the ring position of the substituent is indicated by the number of that atom—for example, 5-methylcytosine, 7-methylguanine, and 5-hydroxymethylcytosine; the element to which the substituent is attached (N, C, O) is not identified. When a chemical group is altered on an exocyclic atom, the type of atom is identified and the ring position to which it is attached is denoted with a superscript. For example, the amino nitrogen attached to C-6 of adenine is N6; the carbonyl oxygen and amino nitrogen at C-6 and C-2 of guanine are O6 and N2, respectively.

RNA Molecules Are Often Site-Specifically Modified In Vivo

Like DNA, many functional RNAs are posttranscriptionally modified at specific nucleotides (Figure 6-34b). Some of the first examples were discovered in ribosomal and transfer RNAs. In some cases, modifications involve the addition of a functional group to an existing nucleotide in the sequence. For example, a methyl group can be added to the 2′ hydroxyl of ribose, thereby blocking its ability to form a hydrogen bond. In bacteria, some C residues of tRNAs are modified to N4-acetylcytidine in a process thought to contribute to the accuracy of protein synthesis. In other cases, the base itself is changed, or its linkage to the ribose—the glycosidic bond—is altered. For instance, inosine, 4-thiouridine, and pseudouridine are relatively common in tRNAs and rRNAs.

Many of the enzymes that catalyze these chemical modifications of RNA have been identified. They are often evolutionarily conserved, indicating that RNA modification has been occurring in biological systems for a long time. More difficult to figure out is the function of these chemical changes in RNA. Molecular biologists can produce unmodified versions of RNAs in the laboratory and compare their functions with those of the chemically altered counterparts isolated from cells. This approach has only rarely discerned much of an effect of a modified base. However, genetic experiments in which an RNA-modifying enzyme is mutated or deleted from an organism suggest that these enzymes give cells a subtle but important selective advantage over organisms that do not modify their RNA. Some evidence supports the hypothesis that RNA modifications stabilize RNA structures and help RNAs interact with proteins in the cell.

The Chemical Synthesis of DNA and RNA Has Been Automated

Knowledge of DNA and RNA chemistry provided the basis for devising methods to synthesize nucleic acids in the laboratory. This technology has paved the way for many biochemical advances that depend on the ability to synthesize oligonucleotides with any chosen sequence. The chemical methods for synthesizing nucleic acids were developed primarily by H. Gobind Khorana and his colleagues in the 1970s. Refinement and automation of these methods have made possible the rapid and accurate synthesis of DNA strands.

DNA (or RNA) synthesis is carried out with the growing strand attached to a solid support (Figure 6-35). First, a nucleotide is attached to the support, a glass or polystyrene bead, through its 3′-hydroxyl group, and polynucleotide synthesis proceeds in the 3′→5′ direction. This is the opposite of the direction of biological polynucleotide synthesis by polymerase enzymes, which is 5′→3′. Functional groups on the bases and phosphates, including hydroxyl and amine groups, are transiently protected with chemical groups that are readily removed after synthesis is complete. The 5′-hydroxyl group is temporarily protected by a dimethoxytrityl (DMT) group; the DMT group is removed from the end of the growing polymer chain at the beginning of each cycle (step 1 in Figure 6-35) to permit extension of the chain by another nucleotide (step 2). Oxidation of the phosphite linkage between the nucleotides completes the cycle (step 3). When chain synthesis is complete, protecting groups are removed from the bases and phosphates, and the oligonucleotide chain is cleaved from its solid support (steps 4, 5, 6). The efficiency of each addition step is very high, allowing the routine laboratory synthesis of polymers containing 70 to 80 nucleotides and, in some laboratories, much longer strands.

Figure 6-35: Solid-phase synthesis of nucleic acids. The oligonucleotide is synthesized in the 3′→5′ direction, starting with a single nucleotide that is covalently attached to a solid support, such as a glass bead (Si). In a repeated series of chemical reactions, nucleotides protected by the dimethoxytrityl (DMT) group from unwanted reactions are sequentially deprotected and reacted to produce a new phosphodiester linkage.

202

Oligonucleotide synthesis is very useful for techniques such as Southern and Northern blotting and for the polymerase chain reaction (PCR) and DNA sequencing, which are discussed in Chapter 7. In addition, chemical synthesis makes it possible to incorporate chemical modifications in the polymer product, such as biotin groups, extra phosphates, sulfhydryl groups, and methyl groups. These functional groups are useful for such applications as specific labeling of a DNA strand or stabilization of an RNA oligonucleotide against enzymatic degradation in cells.

203

SECTION 6.4 SUMMARY

  • Native DNA undergoes reversible unwinding and separation of the strands (melting) on heating or at extremes of pH. DNAs rich in G≡C base pairs have higher melting points than DNAs rich in A = T base pairs.

  • Hybridization, or base pairing between two strands of nucleic acid from different sources, is the basis for important techniques used to study and isolate specific genes and RNAs.

  • Southern blotting is a method by which a specific DNA sequence can be identified in a mixture, following size-based fractionation of the DNA sample by agarose gel electrophoresis. A probe complementary to the DNA of interest is labeled with a radioactive or fluorescent functional group. After transferring the size-fractionated DNA from the gel to a membrane, the probe is hybridized to the sample on the membrane so that the sequence of interest can be visualized.

  • Northern blotting, analogous to Southern blotting, is used for detecting specific RNA sequences.

  • Mutations are alterations in DNA structure that produce permanent changes in the encoded genetic information. Deamination of cytosine is a common chemical mutation in DNA that can damage the genetic code if not corrected by the cell. Deamination of viral nucleic acid can be used to defend against viral infection.

  • Certain A and C residues in DNA are often enzymatically methylated after DNA synthesis. E. coli uses methylation to distinguish between host and foreign DNA and to facilitate the repair of mismatched base pairs that arise from replication errors. In eukaryotes, DNA methylation often inhibits gene expression.

  • Some residues in RNAs are chemically modified by enzymes that introduce methyl or acetyl groups at specific sites or alter a nucleotide base in other ways. These modifications may stabilize RNA structures and can also influence RNA recognition by proteins.

  • DNA and RNA polymers of any sequence can be synthesized with simple, automated procedures involving chemical and enzymatic methods. Solid-phase synthesis of DNA and RNA occurs in the 3′→5′ direction, using chemically protected nucleotides that are selectively deprotected and coupled to the growing polynucleotide chain in successive cycles.

UNANSWERED QUESTIONS

Although many details of nucleic acid structure are well understood, future challenges involve linking the chemistry of these molecules to their behavior in biological systems. Here are several interesting questions in the field.

  1. What are the functions of noncanonical DNA structures in cells? We do not yet know whether non-B-DNA functions in specific cellular processes. For example, some evidence suggests that with its left-handed twist, Z-DNA relieves some of the torsional strain that would otherwise build up during DNA transcription. Perhaps for this reason, the potential to form Z-DNA structures correlates with genomic regions of active transcription. But definitive proof of these ideas has been elusive. Whether three-stranded or four-stranded structures are biologically relevant is also a topic that remains ripe for experimentation.

  2. Do mRNAs have stable three-dimensional structures? Although mRNAs were once thought to be spaghetti-like molecules, increasing evidence hints that they may have stable structures that contribute to biological function. For example, many mRNAs include long sequences that extend beyond the coding region of the gene and are critical for proper gene regulation. Specific proteins bind to these regions and probably recognize structures within them.

  3. How widespread is chemical modification of RNA? Modified nucleotides in tRNA and rRNA have been recognized for a long time, but we do not know whether other RNAs in cells contain such chemical changes. This is an important question, because modifications could influence the function of RNAs that play various roles in controlling gene expression and therefore might be relevant to understanding disease pathways.

204

HOW WE KNOW: DNA Is a Double Helix

Sayre, A. 1975. Rosalind Franklin and DNA. New York: W.W. Norton & Co.

Watson, J.D. 1968. The Double Helix: A Personal Account of the Discovery of the Structure of DNA. New York: Atheneum.

Watson, J.D., and F.H. Crick. 1953. Molecular structure of nucleic acids: A structure for deoxyribose nucleic acid. Nature 171:737–738.

By the early 1950s, DNA had been confirmed as the genetic material in cells, but its structure was unknown. Given that structural information would be key to understanding heredity, the race was on to solve the mystery of DNA structure. Researchers including Rosalind Franklin, Raymond Gosling, and Maurice Wilkins were measuring x-ray diffraction of DNA fibers generated by drawing them from solution using a glass rod. The diffraction patterns reflected the symmetry of the DNA molecules in the fibers, thereby providing an important clue to the molecules’ overall arrangement. By examining the diffraction pattern from fibers oriented perpendicular to the x-ray beam, investigators could deduce the helical symmetry of the molecules inside. These data were interpreted in the light of Chargaff’s rules, which state that in a given DNA sample, the fraction of A equals the fraction of T, and the fraction of C equals the fraction of G. Possible models of the DNA in the fiber were produced, and their calculated diffraction patterns were compared with the experimentally derived patterns.

Rosalind Franklin’s famous Photograph 51 revealed a particularly well-resolved x-ray diffraction pattern of a DNA fiber that was interpreted to determine the 3.4 Å distance between base pairs and the 34 Å periodicity of the helix (characteristic of B-form DNA; see Section 6.2) (Figure 1). The darker spots are areas where the film was hit repeatedly with diffracted x-rays from repeating parts of the DNA molecule. At the top and bottom of the photograph, for example, dark patches represent the nucleotide bases of DNA—the patches are dark because the many bases in the DNA fiber are arranged in a regular pattern. The distance between bases in the DNA structure could be determined by measuring the distance between the dark patches on the film and then making a calculation based on how far the DNA sample was from the x-ray film and how it was positioned relative to the direction of the incident x-ray beam. Watson and Crick made extensive use of this image, along with related diffraction data, to develop a model of the three-dimensional structure of DNA that proved to be correct (Figure 2). This was done at a time before sophisticated computer modeling was possible: Watson and Crick presented their work as a physical model of the double helix constructed on a wire support! Unlike other, competing models of DNA, the Watson-Crick structure had the sugar–phosphate backbone winding around the outside of the helix, with the bases pointing to the interior, where they formed base-pairing interactions between the two strands.

FIGURE 1 Franklin’s “Photograph 51” provided the information necessary to solve the double-helical structure of DNA.
FIGURE 2 A replica of the DNA model built by Watson and Crick. The original model is on display in the Science Museum in London.

Watson and Crick’s work was published in a letter to the British journal Nature in 1953. In the same issue, several other papers provided experimental support for the Watson-Crick model. The double-helical structure immediately suggested a mechanism by which DNA strands could be faithfully copied from one generation to the next. In a famously understated final sentence of their paper, Watson and Crick wrote: “It has not escaped our notice” that the specificity of base pairing could ensure accurate DNA replication.

205

DNA Helices Have Unique Geometries That Depend on Their Sequence

Wing, R., H. Drew, T. Takano, C. Broka, S. Tanaka, K. Itakura, and R.E. Dickerson. 1980. Crystal structure analysis of a complete turn of B-DNA. Nature 287: 755–758.

The discovery of the DNA double helix marked the dawn of molecular biology. However, it was not until 1980 that the first single crystal of a DNA molecule was obtained. This was an important landmark in its own right, because for the first time it became possible to determine the exact helical parameters of a defined DNA sequence. Why did it take almost 30 years after the work of Watson, Crick, Franklin, Wilkins, and Gosling for specific DNA sequences to be crystallized?

The answer is technology. Until the late 1970s, it wasn’t possible to synthesize DNA molecules in the laboratory, so investigators could not produce enough of a specific sequence to make growth of single crystals feasible. Once the methodology was available to synthesize DNA oligonucleotides on solid supports, short DNA molecules of specific length and sequence could be produced in milligram quantities. This material could be purified, and it crystallized readily when concentrated slowly in the presence of suitable buffers. Single crystals of DNA offered some distinct advantages over the DNA fibers analyzed by Franklin, Gosling, and Wilkins. DNA fibers can readily form from a mixture of DNAs of different lengths and sequences, and the structures obtained by analyzing the fiber diffraction patterns produce an “averaged” structure of all the molecules in the fiber. In contrast, single crystals, by definition, are formed by arrays of identical molecules.

Richard Dickerson and his colleagues recognized the wealth of information to be gained by solving a molecular structure of single DNA crystals. They used a self-complementary dodecamer sequence, CGCGAATTCGCG, to solve the first single-crystal structure of DNA. The overall double-helical structure agreed well with that determined by Watson and Crick, but many new details about the geometry of the helix were revealed (Figure 3). This structure, known as the Dickerson dodecamer, ushered in an era of high-resolution structural determinations of DNA and, eventually, the crystallographically determined structures of specific DNAs bound to protein partners. The study of individual DNA sequences also led to extensive studies of DNA–small molecule interactions and to research on the effects of DNA mutations on helical geometry. This work guided the development of certain anticancer drugs, such as cisplatin, that bind and distort DNA and thereby disrupt its replication in rapidly growing cells.

FIGURE 3 The Dickerson dodecamer structure revealed, for the first time, the details of helical geometry for a specific DNA sequence. The drawings are oriented to show the major groove (left) and minor groove (right) in the B-DNA helix.

206

HOW WE KNOW: Ribosomal RNA Sequence Comparisons Provided the First Hints of the Structural Richness of RNA

Gutell, R.R., N. Larsen, and C.R. Woese. 1994. Lessons from an evolving rRNA: 16S and 23S rRNA structures from a comparative perspective. Microbiol. Rev. 58:10–26.

Ribosomes have been around for a long time, and the sequences of the RNAs they contain have been constrained over the course of evolution by the requirements of making functional ribosomes to catalyze protein synthesis. Carl Woese recognized in the 1960s that comparing ribosomal RNA sequences would provide valuable information about the evolutionary relationships among different organisms. Working over many years, he and his colleague Harry Noller assembled careful alignments of the 16S and 23S rRNAs from a large number of microbes. This work led Woese to propose the three-domain theory of life: Eubacteria (now classified simply as Bacteria), Archaebacteria (now Archaea), and Eukarya (eukaryotes).

The comparative analysis approach begun by Woese and Noller was continued by Robin Gutell, who expanded the comparison to include 16S and 23S rRNA sequences from multicellular organisms, including humans. Gutell’s critical analysis provided the first hints that these RNAs form specific three-dimensional structures important to their function. One of the key insights from comparative rRNA sequence analysis was the discovery of noncanonical (i.e., non-Watson-Crick) base pairings. Although these had already been observed in tRNA structures, the much larger sizes of rRNA sequences provided vastly more data. For both 16S and 23S rRNAs, much of the sequence could be folded up into base-paired segments (Figure 4). Comparisons between species showed that the base pairings were much more conserved than were the actual nucleotide sequences. This was because a change in the identity of a nucleotide on one side of a base-paired stretch was typically matched by a mutation in its base-pair partner such that base pairing was maintained. Gutell and Woese also noticed that in many cases, such compensatory base changes occurred for base “pairs” not previously thought to form, such as G–U, A–A, and G–A (see Figure 6-24). In this way, long before high-resolution structures of large RNAs became available, it was clear that RNA molecules are much more tolerant of non-Watson-Crick base pairings than is DNA.

FIGURE 4 This model of a portion of the secondary structure for E. coli 16S rRNA shows canonical base pairs connected by red lines, G–U pairs connected by small black dots, A–G pairs connected by open circles, and other noncanonical pairings connected by solid circles.

207