How DNA Replicates

Matthew Meselson Franklin W. Stahl

hypothesis on dna replication

Matthew Meselson had a passion for physics and chemistry throughout his early life, often conducting science experiments in his family's garage. At the age of 16, he enrolled at the University of Chicago, beginning an academic career that led to doctoral studies at the California Institute of Technology under Linus Pauling. In addition to his widely known work demonstrating semi-conservative replication of DNA with Frank Stahl, Meselson has made many key discoveries in the molecular biology. He is also known for his work in limiting the proliferation of chemical and biological weapons. Meselson is a member of the National Academy of Science and a recipient of the Lasker Award. He continues to serve as a member of the faculty at Harvard University, where he has taught and conducted research since 1960.

hypothesis on dna replication

Following a sheltered life in a Boston suburb (Needham), Frank stumbled his way through college (Harvard, 1951) before fleeing to a graduate school in biology (U Rochester) to avoid the military draft. While in the graduate school, Frank took a course taught by A. H. (Gus) Doermann, and, for the first time in his life, he had a goal. With Gus, he studied genetic recombination in phage. To meet a departmental requirement, Frank took a summer course at Woods Hole, where he met Matt Meselson and began the work described in this Key Experiment. In 1959, Frank joined the faculty at the University of Oregon, Eugene, in their new Institute of Molecular Biology. He has been there ever since. Frank is now an emeritus faculty member who still enjoys teaching as well as family life and the natural wonders of Oregon.

What's the Big Deal?

Some experiments have proven so influential that they have been christened with the names of the scientists who performed them. The "Meselson–Stahl experiment" is one of those. It has also been called "the most beautiful experiment in biology," a title that has seemed to stick over the years. Why was the Meselson and Stahl experiment so important? Their experiment provided the first critical test of the Watson–Crick models for the structure of DNA and its replication, which were not universally accepted at the time. The convincing results of the Meselson–Stahl experiment, however, dispelled all doubts. DNA was no longer just an imaginary model; it was a real molecule, and its replication could be followed in the form of visually compelling bands in an ultracentrifuge. Meselson and Stahl found that these DNA bands behaved in the ultracentrifuge exactly as Watson and Crick postulated they should. Why was the Meselson–Stahl experiment "beautiful"? Because it was conceptually simple and yet sufficiently powerful to differentiate between several competing hypotheses for how DNA might replicate. Taken together, the Watson–Crick model and the Meselson–Stahl experiment marked the transition to the modern era of molecular biology, a turning point as impactful as the theory of evolution. The story of the Meselson–Stahl experiment, as told here by its protagonists, also reveals how friendship and overcoming obstacles are as crucial to the scientific process as ideas themselves.

Learning Overview —

Big concepts.

Faithful replication of the genetic material (DNA) is the foundation of all life on earth. The experiment by Meselson and Stahl established that DNA replicates through a semi-conservative mechanism, as predicted by Watson and Crick, in which each strand of the double helix acts as a template for a new strand with which it remains associated, until the next replication.

Bio-Dictionary Terms Used

Bacteriophage (phage) , base , base pairing , chromosome , DNA , Hershey–Chase experiment , eukaryote , mutation , nucleotides , Prokaryote (bacteria) , recombination , RNA , ultraviolet light

Terms and Concepts Explained

Equilibrium density-gradient centrifugation , DNA replication , isotope , semi-conservative DNA replication

Introduction

Matthew Meselson and Franklin Stahl (both 24 years old) met at the Marine Biological Laboratory in Woods Hole in Massachusetts and decided to test the Watson–Crick model for DNA replication, which was unproven at the time.

What Events Preceded the Experiment?

Watson and Crick proposed a "Semi-Conservative" model for DNA replication in 1953, which derived from their model of the DNA double helix. In this proposal, the strands of the duplex separate and each strand serves as a template for the synthesis of a new complementary strand. Watson's and Crick's idea for DNA replication was a model, and they did not have data to support it. Some prominent scientists had doubts.

Two other models, "Conservative" and "Dispersive", for DNA replication were proposed.

Setting Up the Experiment

A method was needed to detect a difference between the parental and daughter (newly replicated) DNA strands. Then, one could follow the parent DNA molecule in the progeny. Meselson thought to distinguish between parental and newly synthesized DNA using a density difference in the building blocks (nucleotides) used to construct the DNA. The three models for DNA replication would predict different outcomes for the density of the replicated DNA in the first- and second-generation daughter cells.

The general experimental idea was first to grow bacteria in a chemical medium to make high-density DNA and then abruptly shift the bacteria to a low-density medium so that the bacteria would now synthesize lower density DNA during upcoming rounds of replication. The old and newly synthesized DNA would be distinguished by their density.

To measure a density difference in the DNA, Meselson and Stahl invented a method called equilibrium density gradient centrifugation. In this method, the DNA is centrifuged in a tube with a solution of cesium chloride. When centrifuged, the cesium chloride, being denser than water, forms a density gradient, reaching a stable equilibrium after a few hours. The DNA migrates to a point in the gradient where its density matches the density of the CsCl solution. Heavy and light DNA would come to different resting points and thus physically separated.

Doing the Key Experiment

Meselson and Stahl first decided to study the replication of DNA from a bacteriophage, a virus that replicates inside of bacteria, and used a density difference between two forms of the nucleobase thymine (normal thymine and 5-bromouracil). These experiments did not work.

The investigators changed their plans. They studied replication of the bacterial genome and used two isotopes of nitrogen (15N (heavy) and 14N (light)) to mark the parental and newly synthesized DNA.

When the population of bacteria doubled, Meselson and Stahl noted that the DNA was of an intermediate density, half-way between the dense and light DNA in the gradient. After two doublings, half of the DNA was fully light and the other half was of intermediate density. These results were predicted by the Semi-Conservative Model and are inconsistent with the Conservative and Dispersive Models.

Meselson and Stahl did another experiment in which they used heat to separate the two strands of the daughter DNA after one round of replication. They found that one strand was all heavy DNA and the other all light. This result was consistent with the Semi-Conservative model and provided additional evidence against the Dispersive Model.

Overall, the results provided proof of Semi-Conservative replication, consistent with the model proposed by Watson and Crick.

What Happened Next?

Within a couple of weeks after their key experiment, Meselson wrote a letter to Jim Watson to share news of their result (letter included).

Max Delbruck, the Caltech physicist and biologist who had proposed the dispersive model, was elated by the results, even though Meselson and Stahl disproved his replication hypothesis, and urged the young scientists to write up their results for publication and announce the important result to the world (1958).

Scientists now know a great deal about the protein machinery responsible for DNA replication.

Closing Thoughts

The Meselson–Stahl experiment had a powerful psychological effect on the field of genetics and molecular biology. It was the first experimental test of the Watson and Crick model, and the results clearly showed that DNA was behaving in cells exactly as Watson and Crick predicted.

In addition to having a good idea, the behind-the-scenes tour of the Meselson–Stahl experiment reveals that friendship and persistence in overcoming initial failures play important roles in the scientific discovery process. Also important was an atmosphere of freedom that allowed Meselson and Stahl, then very junior, to pursue their own ideas.

Guided Paper

Meselson, M. and Stahl, F.W. (1958). The replication of DNA in Escherichia coli. Proceedings of the National Academy of Sciences U.S.A., 44: 672–682.

This classic paper provides experimental evidence that the strands of the DNA double helix serve as templates to create a new copy of DNA. These results provide experimental evidence of The Watson and Crick model of DNA replication (‘semi-conservative replication’) demonstrating that genetic information is passed from one cell or organism to its progeny.

The first conversation between Matt Meselson and Frank Stahl, in the summer of 1954, began a collaboration that led to their Key Experiment on DNA replication and marked the beginning of a lifelong friendship. Matt and Frank describe the circumstances that brought them together below.

It was 1954, the year after Jim Watson and Francis Crick published their two great papers describing their double helical model of DNA and its implications for how it might replicate, mutate, and carry genetic information. Jim Watson (26 years old) and I (a 24-year-old first-year graduate student) were both at Caltech and living at the Athenaeum, the Caltech faculty club. We often talked while waiting for dinner. One day, Jim asked me to join him for the summer as his teaching assistant in the Physiology Course at the Marine Biological Laboratory (MBL) at Woods Hole, Massachusetts. He mainly wanted me to do experiments to see if RNA was a double helix. (As a side note, those experiments showed that RNA is not a double helix.) So in June 1954, I drove my 1941 Chevrolet coupe across the country from Cal Tech in Pasadena, California, to the MBL at Woods Hole, Massachusetts.

One day in Woods Hole, while planning student assignments for the Physiology Course, Jim went to the window of the course office upstairs in the MBL Lillie building and pointed towards a student sitting on the grass under a tree serving gin and tonics. That student was Frank Stahl. Jim said let's give him a really hard experiment to do all by himself – the Hershey–Chase Experiment; then, we'll see how good he really is. Two years earlier, Alfred Hershey and Martha Chase published an influential experiment that provided evidence implicating DNA, and not protein, as the substance conferring genetic inheritance in bacteriophage (see the whiteboard animation video on the Hershey–Chase experiment ).

I thought that this guy serving gin and tonics must be an interesting fellow, so I went downstairs to meet him and let him know what was being planned for him. Frank was then a biology graduate student at the University of Rochester. I sat down on the grass under the tree, and we hit it off right away. We found that we had much to discuss. I was very impressed with Frank's knowledge of phage genetics, a subject I knew nothing about, coming from the laboratory of Linus Pauling where I was doing X-ray crystallography. Frank can tell you more about our conversation under the "gin and tonic tree."

1954 was an exciting time for molecular biology. One year earlier, Watson and Crick published their model for the double helix structure of DNA (see the Narrative on DNA Structure by Vale ), which aroused much excitement as well as some serious disbelief. With Watson as an instructor and Crick and Sydney Brenner as visitors and several other founders of molecular biology, Woods Hole that summer became an epicenter for discussion of the great questions in molecular biology. Could the double helix model, as Watson and Crick proposed, explain the replication of the genetic information? What is the "code" for reading out the nucleotide sequence of DNA and turning that into the sequence of amino acids in protein? Would RNA have a similar structure to DNA? However, at the time of my arrival, I had no idea that Woods Hole was hosting anyone, but me, interested in the big problems of modern, i.e., molecular, biology.

I was a 24-year-old biology graduate student at the University of Rochester and had come to the MBL to take the Physiology Course. To beat the heat and, perchance, to meet someone interesting, I invested in a bottle of gin and a 6-pack of tonic, found some ice, a thermos jug, and a tree and sat myself down in the shade. Matt Meselson was one of my first catches.

Our conversation under the "gin and tonic tree" was life-changing. After Matt warned me of Watson's planned test of my experimental aptitude, we talked about the work we were doing. I explained to him the problem in phage genetics on which I was working. Eventually our conversation turned to DNA replication. Neither of us was working on the problem at the time, although we were both keenly interested in it. It was perhaps the most important contemporary question in molecular biology.

At Caltech, Matt had already come up with an idea for how the mechanism of DNA replication might be studied by density labeling. But, as a physical chemist, he was unfamiliar with the methods of phage and bacterial biology that would be needed to conduct the actual experiments. So we decided to collaborate. I was planning to be a postdoctoral fellow at Caltech starting that September. If we could develop a method for measuring the density of DNA molecules and successfully apply it to the problem of DNA replication, we could establish whether the Watson–Crick model for DNA replication, and even the model of the structure itself, was correct or not.

We did not begin our collaboration immediately because Matt needed to finish his X-ray crystallography and I had made plans to do my postdoctoral work on bacterial genetics with Joe Bertani.

You can also hear Matt Meselson describing the Meselson–Stahl experiment in Video 1 .

The genetic material of eukaryotic cells is organized in the form of chromosomes , each a single linear, double-stranded DNA molecule ( Figure 1 ). Most prokaryotes (bacteria) have a single, circular chromosome ( Figure 1 ). All forms of life must replicate their DNA and, except for recombination and infrequent mutations, pass identical copies of their genetic material to their progeny. (See also the Whiteboard Video on Keeping Track of Your DNA .)

hypothesis on dna replication

Based upon their model for the structure of DNA, Watson and Crick proposed that DNA replicates in a "Semi-Conservative" manner ( Figure 2 ). In this model, the two strands of the DNA double helix unwind and separate, and each "parent" strand serves as a template for the synthesis of a new "daughter" strand. The Watson–Crick base pairing (see the Narrative on DNA Structure by Vale ) of adenine with thymine and guanine with cytosine ensures that, except for rare copying errors, mutations, the information of the original double-stranded DNA molecule is preserved during the synthesis of the daughter strands. In the Semi-Conservative model, each daughter cell in the first generation would inherit one of the original DNA strands from the parent and a recently synthesized DNA strand. In the second generation, two of the granddaughters would be composed of all newly synthesized DNA and two granddaughters would have hybrid DNA (one parental strand and one newly synthesized strand).

hypothesis on dna replication

While (spoiler alert) the Semi-Conservative Model turned out to be correct, it was far from a foregone conclusion. Before our experiment, several leading scientists questioned the Semi-Conservative Model (see Dig Deeper 1 for more information on their reservations) and proposed alternate models discussed below.

Explorer's Question: What do you imagine are the pros and cons of this model?

Answer: The beauty of this model is that it provides a clear explanation of how a daughter strand is made from the template strand ( Figure 3 ). However, the model requires that the two long parental DNA strands unwind to become single-strand templates. This was seen as a weakness by many scientists at the time (see Dig Deeper 1 for more information).

hypothesis on dna replication

The unwinding of the DNA helix and keeping the daughter and parental strands from becoming tangled posed problems for the Semi-Conservative model in the minds of many scientists. As a result, other models for DNA replication were imagined. One was a "Conservative Replication" model ( Figure 4 ). In this model, the parental double helix forms a template for a completely new double helix. The two original strands remain together, no unwinding occurs, and the daughter DNA is formed from newly synthesized DNA. In this model, in the first generation, one daughter DNA would inherit the original DNA double helix from the parent DNA and the other daughter DNA would inherit the newly synthesized DNA double helix. In the second generation, one of the four granddaughters would have the original parental DNA and the other three granddaughters would all be composed of newly synthesized DNA. While Conservative Replication was a logical possibility, it was not elaborated by any specific mechanism.

hypothesis on dna replication

Answer: This model did not call for unwinding of the DNA strands, as in the Semi-Conservative Model, thus solving the concern about DNA unwinding. However, it was unclear how the copying machinery would read out the nucleotide sequence information buried in the core of the DNA double helix.

A third possibility was a model proposed by Max Delbrück (later called "Dispersive Replication") ( Figure 5 ). Delbrück doubted that the two strands of the double helix could be unwound or pulled apart to undergo Semi-Conservative replication and instead suggested that DNA strands broke at every half-turn of the helix during replication (discussed in more depth below and in Dig Deeper 1 ). According to Delbrück's Dispersive Replication Model, each helix of the replicated DNA consists of alternating parental and daughter DNA. Unlike the other two models, the progeny in subsequent generations would be indistinguishable with regard to their compositions of parental and newly synthesized DNA.

hypothesis on dna replication

Answer: Fragmentation would create shorter templates for replication, which would minimize any unwinding or untangling problems faced by one very long DNA molecule. However, the reassembly of the fragments again into the intact chromosome could be problematic, especially if the breaks occur at every half-turn of the helix.

Explorer's Question: In the first generation, which model(s) would predict that the two daughter cells would receive approximately equal amounts of the original and newly synthesized DNA?

Answer: The Semi-Conservative Model and the Dispersive Model. However, differences in daughter composition arise in the second generation in the two models.

Matt and Frank learned about the models for DNA replication prior to their first meeting at the Physiology Course at the Marine Biological Laboratory through circumstances described below.

Sometime in 1953, while I was a graduate student of the great chemist Linus Pauling, I went to see Max Delbrück, a physicist and founder of the "phage group" who had become deeply interested in genetics and the basis of life (Max Delbrück, and his work with Salvador Luria, is featured in the Narrative on Mutations by Koshland ). I wanted to learn what problems in biology he thought were most important and what advice he might have for me about getting into biology. Almost as soon as I sat down in his office, he asked what I thought about the two papers by Watson and Crick that had been published in Nature earlier that year. I confessed that I had never heard of them.

Exasperated, Delbrück flung a little heap of reprints of the Watson–Crick papers at me and shouted "Get out and don't come back until you have read them." What I heard was "come back." So I did, but only after reading the papers.

When I came back, Delbrück said he did not believe in the Semi-Conservative mechanism of DNA replication proposed by Watson and Crick. Max had imagined that if replication is semi-conservative, the two daughter duplexes would become wound around each other turn-for-turn as the two chains of the mother molecule became unwound. To get around the supposed problem of untangling the daughter molecules, he proposed a model in which breaks are made to prevent interlocking when separating, and then joined back together (see Figure DD1 in Dig Deeper 1 ). This mechanism required rotation of only short lengths of duplex DNA, after which the chains would be rejoined. In the rejoining process, sections of the new chain would be fused to sections of the old chain, making all four of the chains mosaics of new and old DNA. Delbrück, in 1954, published a paper that questioned the Watson–Crick model and presented this new model (later referred to as "Dispersive Replication" as shown in Figure 5 and Dig Deeper 1 ). In some ways, the idea of Delbrück was ahead of its time. Transient breaks are now known to be made by an enzyme called topoisomerase, but in a manner that leaves the individual chains of the parent duplex intact (see Dig Deeper 1 ).

In addition to Max's reservations and model, several other scientists also posed their own concerns and solutions to the "unwinding problem" or alternatives to the Watson–Crick DNA structure itself (see Dig Deeper 1 ).

What I gathered from my conversations with Max and others was that not everyone believed the DNA replication model of Watson and Crick. It was only a hypothesis with no experimental evidence to support it. The key to solve this problem was to follow the parental DNA in the progeny. But how?

I was working on something very different for my PhD thesis with Linus Pauling, but earlier that year, I had an idea for labeling protein molecules with deuterium, a heavy isotope (2H) of hydrogen (1H) and separating them from unlabeled protein molecules in a centrifuge according to their density as a means to solve a quite different problem (see Dig Deeper 2 ). After that second meeting with Max, it occurred to me that density labeling and centrifugal separation might be used to solve the DNA replication problem. When I told this to Pauling, he urged me to get my X-ray crystallography done first. And when I proposed the density approach to Watson, one of those evenings waiting for dinner at the Athenaeum, he said I should go to Sweden to do it – where the ultracentrifuge had been invented (which I never did).

My entry point to thinking about DNA replication came when I was trying to understand how bacteriophage (viruses that infect bacteria) exchange pieces of DNA with one another. This process of DNA exchange between chromosomes is called recombination (see the whiteboard video on the experiments by Morgan and Sturtevant ). When did this recombination process occur? Did it occur when DNA replicates? Or perhaps recombination was an event that stimulated DNA replication? My intuition was that the processes of recombination and replication were somehow related. However, little was known about the mechanisms of either DNA replication or recombination at the time. Furthermore, I did not know how to pursue these questions in 1954. My ideas for experiments were lame and would have led to un-interpretable data.

Like many interesting questions in biology, often one has to be patient until either the right idea or technology emerges that allows one to answer them properly. In 1954, my awareness of a possible connection between replication and recombination primed my interest for the first gin and tonic conversation with Matt. However, it was several decades before I was sure that, in bacteriophage, DNA replication and recombination, in a large degree, depended upon each other. The convincing experiments were based on variations of a technique pioneered by Matt and Jean Weigle at Caltech. In this method, density-labeled, genetically marked parental phage infect the same bacteria. The densities and genetic makeup of progeny phage are determined by bioassay of the individual drops collected from a density gradient.

Matt and Frank

To distinguish between the Semi-Conservative, Conservative, and Dispersive Models of DNA replication described above, we needed a method that could tell the difference between the parental and daughter DNA strands. Figures 2 , 4 , and 5 illustrate the parent and newly synthesized strands with different colors. However, we needed to find a real physical difference that would serve the same function of distinguishing between the old and newly synthesized DNA. Matt had the idea of distinguishing old and new DNA by having the bacteria synthesize them with different isotopes and separating them in a centrifuge according to their density. If the original and the newly synthesized DNAs could be made of different density materials, then we could perhaps measure this physical difference. We will discuss the chemicals that were used to make the DNA heavier or lighter in the next section.

Our experimental idea was to grow an organism in a chemical medium that would make its DNA heavy. Then, while it was growing, we would switch to a new medium in which the newly synthesized DNA would be made of lighter material ( Figure 6 ). The density difference between the original and the newly replicated DNA could allow us to distinguish between models for DNA replication.

hypothesis on dna replication

To separate DNA of different density, we invented a method, called "equilibrium density gradient centrifugation," and published it, together with Jerome Vinograd, a Caltech Senior Research Fellow who had taught us how to use the ultracentrifuge in his lab and provided advice. In this method, as applied to DNA, a special tube that has quartz windows so that ultraviolet light photos can be taken while the centrifuge is running is filled with a solution of cesium chloride and the DNA to be examined. Upon centrifugation at high speed (~45,000 revolutions per minute or 140,000 times gravity), the CsCl gradually forms a density gradient, becoming most concentrated at the bottom of the tube ( Figure 7 ). The CsCl solution toward the top of the tube is less dense than the DNA, while the CsCl solution at the bottom is denser than DNA. Thus, when a mixture of DNA in a CsCl solution is centrifuged, the DNA will eventually come to a resting point where its density matches that of the CsCl solution ( Figure 7 ). The DNA absorbs UV light, and its position along the tube was recorded by using a special camera while the centrifuge is running.

hypothesis on dna replication

The method now seems straightforward, but in reality, it took a couple of years to develop. For example, we did not come to cesium chloride immediately. We looked at a periodic table for a dense monovalent atom that would not react with DNA; rubidium chloride (molecular weight of 121) was available in the Chemistry Department stockroom and we initially tried to use that but found that even concentrated solutions were not dense enough to float DNA to a banding point. We then moved one level down in the periodic table to cesium (the molecular weight of cesium chloride is 168) and that worked (for more details, see Dig Deeper 3 ).

Frank and Matt

In the fall of 1954, we were reunited in Cal Tech and lived for about eight months in the same house across the street from the lab. We finally could begin doing experiments to test models of DNA replication. It should be noted that DNA replication was our "side" project; we also had our "main" projects under the supervision of our respective professors. However, faculty at Cal Tech was kind in allowing two young scientists to venture forward with their own ideas.

While the general experimental approach that took form under the "gin and tonic tree" was straightforward, choices had to be made in how exactly to do the experiment. What organism should we use? Would a chemical trick of making DNA heavier or lighter work and could we measure a small density difference between the two? It took us a while to get the conditions right, about two years.

We first decided to examine the replication of the bacteriophage T4 inside of the bacterium Escherichia coli. Bacteriophages are viruses that invade and replicate inside of a bacterium; when new viruses are made, they will burst the bacterium and then spread to new hosts. Bacteriophage have small genomes and are therefore the smallest replicating systems. Frank's PhD thesis was on T4, so he knew how to work with this phage. Max Delbrück and others at Cal Tech were also actively studying phage (see the Narrative on Mutations by Koshland ). Thus, T4 seemed the obvious choice. To create DNA of heavier density than normal DNA, we decided to use the analogue, 5-bromouracil, of the base thymine, in which a heavier bromine atom replaces a lighter hydrogen atom. During replication, 5-bromouracil could be incorporated into DNA, instead of thymine.

However, while this approach seemed reasonable, it did not work in practice. Although we did not appreciate it at the time, during phage growth, the DNA molecules undergo recombination, joining parental DNA to newly synthesized DNA in a manner that after several generations would give no clear-cut distribution of old DNA among progeny molecules. Also, we learned from a recent paper that 5-bromouracil was mutagenic and made a detour into studying mutagenesis before coming back to our main project.

We clearly needed a new strategy.

Instead of phage, we decided to study the replication of the bacterial genome. This was a good choice – the bacterial DNA gave a very sharp and clear band when centrifuged in a solution of cesium chloride (to learn more about why we used cesium chloride to create a density gradient and the general use of this technique; see Dig Deeper 3 ).

We also switched our density label. DNA is made up of several elements – carbon, nitrogen, oxygen, phosphorus, and hydrogen. Some of these elements come in different stable isotopes, with atomic variations of molecular weight based upon different numbers of neutrons. Nitrogen-15 (15N) is a heavier isotope of nitrogen (the most common isotope, 14N, has a molecular weight of 14 Daltons). We could easily buy 15N in the form of ammonium chloride (15NH4Cl), which was the only source of nitrogen in our growth medium. The 15N in the medium then found its way into the bacterial DNA (as well as other molecules) in a harmless manner and did not impair bacterial growth.

We also had good luck in that Caltech bought a brand new type of ultracentrifuge called an analytical ultracentrifuge (Model E) developed by the Beckman Instrument Company. The Model E was a massive machine about the size of a small delivery truck (the current model is just a bit bigger than a dishwasher). Importantly, the Model E could shine a UV light beam on the tube while the centrifuge was spinning and detect and photograph the position of the DNA. The good news was that 15N-containing DNA and 14N-containing DNA could be clearly distinguished by their different density positions ( Figure 8 ).

hypothesis on dna replication

Finally, we had everything in place to try our experiment properly. I decided to set up our first experiment in the following two ways:

1) Grow the bacteria in "light" nitrogen medium and then switch to "heavy."

2) Grow another culture of bacteria in "heavy" nitrogen for many generations and then switch to "light."

Frank was called to a job interview and could not perform this first experiment with me. But before leaving, he warned me – "Don't do the experiment in such a complicated way on your first try. You might mix up the tubes."

I ignored Frank's advice and did the experiment both ways.

In the first experiment after transferring bacteria grown in heavy nitrogen (15N) growth medium and then switched to "light" (14N) nitrogen medium, I saw three discrete bands corresponding to old, hybrid, and new DNA, as predicted by Semi-Conservative replication. Excited developing the photograph in the darkroom, I remember letting out a yelp that caused a young woman working nearby to leave in a hurry. But later I realized my mistake. Frank had been prophetic. I indeed had mixed tubes, combining two different samples, one taken before and the other taken after the first generation of bacterial growth in the light medium. As described for the correct experiment below, there is no time when old, hybrid, and fully new DNA are present at the same time.

When I came back from my trip, Matt and I performed what proved to be the decisive experiment. We grew bacteria in "heavy" nitrogen (15N; from 15NH4Cl) and then switched to "light" nitrogen (14N; 14NH4Cl) and, at different time points, collected the bacteria by centrifugation, added detergent to release the DNA, and combined this with concentrated CsCl solution to reach the desired density. After 20 hours of centrifugation and the final density positions of the DNAs had come into view, we knew that we had a clean answer ( Figure 9 ). The DNA from bacteria grown in heavy nitrogen formed a single band in the gradient. However, when the bacteria were shifted to a light nitrogen medium and then allowed to replicate their DNA and divide once (first generation), essentially all DNA had shifted to a new, "intermediate" density position in the gradient ( Figure 9 ). This intermediate position was half-way between the all heavy and all light DNA. At longer times of incubation in light nitrogen, after the cells had divided a second time (second generation), a DNA band at lighter density was seen and there were equal amounts of the intermediate and light DNA.

hypothesis on dna replication

Explorer's Question: Which of the three models (Conservative, Semi-Conservative, or Dispersive) is most consistent with the results of this experiment?

Answer: The Semi-Conservative model. The Conservative model predicts a heavy and light band at the first generation, not an intermediate band. The Dispersive model predicts a single intermediate band at both the first and second generations (the band shifting toward lighter densities with more generation times).

Explorer's Question: Why are the two DNA bands at the 1.9 generation time point of approximately equal intensity?

Answer: After the first generation, each of the two heavy strands is partnered with a light strand. The bacterial DNA consists of one heavy strand and one light strand. When that heavy–light DNA replicates again in the light medium, the heavy strands are partnered with new light strands (intermediate density DNA) and the light strands are also partners with new light strands (creating all light density DNA).

The experiment that Frank described above took hardly any time at all (2 days) and yielded a clean result. We then repeated it without any problem. Once we knew how to set up the experiment, it was relatively easy. But it took us two years of trials before we got the experimental design and conditions right for the final ideal experiment.

The experiment clearly supported the Semi-Conservative Replication model for replication and, in doing so, also supported the double helical model of DNA itself. However, we wanted to do one more experiment that would examine whether the "intermediate" density band of DNA in the first generation was truly made of two and just two distinct subunits, as predicted by the Watson–Crick model. The model predicts that one complete strand of DNA is from the parent and should be heavy and the other complete DNA strand should be all newly replicated and therefore light ( Figure 10 ). We could test this hypothesis by separating the subunits with heat and then analyzing the density and molecular weight of the separated subunits by equilibrium density-gradient ultracentrifugation.

hypothesis on dna replication

On the other hand, the Dispersive Model predicted that each DNA strand of first generation is an equal mixture of original and newly replicated DNAs ( Figure 11 ).

hypothesis on dna replication

The results from the experiment were again clear ( Figure 12 ). The "intermediate density" DNA in the first generation split apart into a light and heavy component. From the width of the DNA band in the gradient (see Dig Deeper 3 ), we could also tell that the light and heavy DNA obtained after heating had each half of the molecular weight of the intermediate density DNA before heating. These results indicated that each parental strand remained intact during replication and produced a complete replica copy. This was decisive evidence against the Delbrück model for it predicted that both strands would be mosaics of heavy and light, not purely heavy and purely light. And the finding that the separated heavy and light subunits each had half the molecular weight of the intact molecule indicated that DNA was made up of two chains, as predicted by the Watson Crick model, and was not some multichain entity.

hypothesis on dna replication

Based upon the results in Figure 9 and Figure 12 , we concluded that:

1) The nitrogen of a DNA molecule is divided equally between two subunits. The subunits remain intact through many generations.

2) Following replication, each daughter molecule receives one parental subunit and one newly synthesized subunit.

3) The replicative act results in a molecular doubling.

These conclusions precisely aligned with the Watson–Crick Semi-Conservative model for DNA replication. DNA, as a double-stranded helix, unwinds, and each strand serves as a template for the synthesis of a new strand.

When we had our result, Matt quickly shared the news with Jim Watson in a letter dated November 8, 1957 (available for the first time here ). It was common in those days to share results with colleagues through letters prior to a publication.

We also shared our results with Max Delbrück who took the news well that his Dispersive Replication model was incorrect. In fact, he wrote to a colleague that Meselson and Stahl had obtained a "world shaking result." But we were slow to get our work written up for publication. Once we knew the answer, we were keen to move onto new experiments rather than writing up our results.

Finally, Max had enough of our dallying and brought us down to the Caltech marine station at Corona del Mar. There, he quarantined us to a room in a tower, saying that we could not come out until we had written a draft of our paper. He was not being unkind, and we thought it great fun. Max's wife Manny Delbrück kindly came in occasionally to bring us delicious sandwiches, and Max also kept us company. We worked for 2 days straight and got him a draft.

Shortly thereafter, we completed our paper and Max communicated it in May 1958 to the Proceedings of the National Academies of Science, 4 years after our meeting at the Marine Biological Laboratory but less than a year after finally getting our experiments to work.

After our paper was published, we went separate ways in our lives. Frank got a job at the University of Missouri but soon thereafter moved to the University of Oregon in Eugene. Matt got promoted from a postdoctoral fellow to an assistant professor at Caltech and was teaching physical chemistry. However, the constant teaching limited time in the laboratory. Matt asked to be demoted from assistant professor back to senior postdoc, so he could get more work done in the lab. This is perhaps the only case in the history of Caltech in which a professor asked to move down the academic ladder. After a year as a senior postdoc, Matt then moved to Harvard to become an associate professor.

Decades have passed, and we now know much about the machinery that orchestrates DNA replication, including the unwinding of the strands and the synthesis of a new strand from the parental template. The details are beyond what can be discussed here, but you can view an animation of this process in Video 2 .

When we first discussed the use of density labeling to test models for DNA replication under the gin and tonic tree, we could not imagine the psychological effect our experiment would have on the field. Many scientists were not initially convinced by the Watson–Crick model for the structure of DNA or their proposal for its mode of replication. It was not clear if their model could explain heredity and the properties of genes. Some people seemed to think the model was too simple to be the gene. Others thought it too simple (meaning too beautiful) to be wrong!

However, after our experiment, the DNA model seemed very real. We could watch DNA with a camera; the visualization of DNA bands was simple and clear. Our results showed that the gene is made of two complementary halves, each a template for the other. Even the disbelievers, such as the deeply thoughtful Max Delbrück, acquiesced. DNA was no longer an imaginary molecule in the heads of Watson and Crick. It was a dynamic molecule; one could perform experiments on it, and it behaved in living cells as one might predict. Mendel's concept of a discrete "factor" that could determine a plant character and remain intact generation after generation and the physical reality of a gene as double-stranded DNA became intertwined from that moment on.

It is gratifying to think that our experiment, so simple by modern standards, is still valued and taught. But beyond the logic of how the experiment was performed, we hope that our story also conveys other important lessons about science.

• Every hypothesis needs to be rigorously tested with a clear experiment.

• An atmosphere of freedom is important. We were both very junior at the time of this experiment, but we were supported by senior scientists who encouraged us to pursue our own ideas.

• Success does not come immediately. Reading most scientific papers (including ours), everything seems straightforward and works right away. Our narrative shows that the so-called "most beautiful experiment in biology" had some unsuccessful excursions and two years of work to come to successful finish.

• Because success does not come immediately, it is valuable to be able to share difficult times with a friend. We kept each from getting discouraged. There was a certain gaiety in our work. We even had fun when things went wrong.

• Much of science is built upon collaboration and friendship. This Key Experiment could never have been the "Meselson Experiment" or the "Stahl Experiment." The "Meselson and Stahl Experiment" required both parties. We complemented each other scientifically and encouraged each other personally. Well more than a half-century has passed since this experiment was performed, and we remain good friends today.

Dig Deeper 1: Alternatives to Semi-Conservative replication

Max Delbrück, in his 1954 paper (PNAS 40: 783-788), said the following of the Watson and Crick Semi-Conservative replication mechanism:

"The principal difficulty of this mechanism lies in the fact that the two chains are wound around each other in a large number of turns and that, therefore, the daughter duplexes generated by the process just outlined are wound around each other with an equally large number of turns. There are three ways of separating the daughter duplexes: (a) by slipping them past each other longitudinally; (b) by unwinding the two duplexes from each other; (c) by breaks and reunions. We reject the first two possibilities as too inelegant to be efficient and propose to analyze the third possibility."

Max's solution was to break the single chains at regular intervals, allowing rotation about single bonds of the unbroken chain followed by joining in a way that dispersed short segments of parental DNA among the single chains of the daughter duplexes. This is the Dispersive Model presented in Figure 5 and presented in more detail in Figure DD1 . There was a germ of truth in Delbrück's idea of breakage. We now know of topoisomerases, enzymes that facilitate DNA replication by temporarily breaking, allowing unwinding, and then rejoining DNA. There are also enzymes that unwind DNA helixes called DNA helicases. Both enzymes use chemical energy derived from hydrolyzing adenosine triphosphate to perform work on the stable DNA double helix.

hypothesis on dna replication

Another type of solution to the "unwinding problem," one that required no breaking and no entangling of the daughter molecules, was to imagine that the synthesis process would cause the entire parental duplex to rotate one turn for each turn of DNA synthesized. But this posed problems of its own – giving rise to a variety of long-forgotten proposed models, including evoking a motor at the growing point that would drive the rotation of the parent molecule, as proposed by John Cairns and Cedric Davern [J. Cellular Physiology, 70: S65–76 (1967)].

Alternative solutions questioned the DNA double helix model, but not semi-conservative replication. For example, one idea was to assume that the two chains are not wound around a common axis, but instead are simply pushed together (plectonemic coiling), which would require no unwinding and no rotation. This possibility, although it appeared remote, was not rigorously ruled out until much later in a paper by Crick, Wang, and Bauer in 1979 (J. Mol. Biol, 129: 449–461).

Dig Deeper 2: The idea for using density for separation

The idea for using density as a separation method came to me early in 1954 while I was a first-year graduate student at Caltech listening to a lecture by the great French scientist Jacques Monod. Monod was describing the problem of regulation of an enzyme called beta-galactosidase. If the bacteria were growing in a medium without lactose (a sugar), the enzyme activity was very low. When lactose or a chemical analogue of lactose was added, the enzyme activity was induced. The question was how? One model was that the enzyme was always there, but is inactive unless lactose is around. Another model (the correct one) was that the enzyme is synthesized de novo after the inducer is added. I thought that it might be possible to measure new enzyme and distinguish it from old enzyme if the enzyme was synthesized from heavier building blocks (amino acids). How could one make heavier building blocks? I thought that deuterium (a heavy isotope of hydrogen; 2 H) might be the answer. If one grew bacteria in heavy water ( 2 H 2 O) and switched to normal water ( 1 H 2 O) when one added inducer, then any newly synthesized beta-galactosidase would have had a greater density than the pre-existing beta-galactosidase. I never did the experiment, but the idea primed me for the DNA replication problem.

Dig Deeper 3: The role of the centrifuge in the Meselson–Stahl experiment

Matt describes briefly how this technique evolved

The first paper (see the reference list) that Frank and I wrote together (along with Jerome Vinograd) was on the method and theory of using centrifugation in an equilibrium density gradient, which showed that this method not only could separate molecules but could also be used as a tool to determine their molecular weights. This work was also part of my PhD thesis at Caltech. When I presented this work at my thesis defense, the great physicist Richard Feynman was on the examination committee, along with Pauling, Vinograd, and one of Pauling's post-docs who taught me X-ray crystallography. Feynman had not read the thesis but did so during the defense. I presented my rather long mathematical derivation showing that macromolecules in a density gradient in a centrifugal field would be distributed in a Gaussian manner about the position of neutral buoyancy with the width dependent upon the square root of the molecular weight. Feynman then went to the blackboard and, on the spot, produced a much shorter derivation of the same thing, modeled on the wave function for the quantum mechanical harmonic oscillator. Feynman writes about our experiment in his jolly book, Surely You're Joking, Mr. Feynman .

The way in which we found that CsCl forms a density gradient on its own was somewhat fortuitous. We initially thought that we needed to pour a CsCl gradient in the tube in advance. However, we found that just by centrifuging an initially homogeneous solution of cesium chloride produced a continuous gradient density on its own after several hours. From the width of the DNA band in the density gradient, we could also calculate the molecular weight of the DNA molecules in the gradient as 7 million Daltons. The chromosome of E. coli is much, much larger, but long molecules of DNA are fragile and had been broken up by shear forces while passing through the hypodermic needle with which we loaded the centrifuge cell. Subsequent to our result, CsCl equilibrium density-gradient centrifugation became a standard tool for isolating DNA from cells for decades and was used in important experiments such as the demonstration of messenger RNA by Brenner, Meselson, and Jacob and showing the mechanism of general recombination in phage lambda by Frank Stahl.

References and Resources

  • Matthew Meselson’s letter to James Watson from November 8, 1957, describing the results of their experiments on DNA replication. Download .

This paper describes the use of the centrifuge and density gradient to analyze biological molecules, a technique that was used in their 1958 paper but also very broadly used for many applications in biology. See also Dig Deeper 3 .

An outstanding resource for those wanting a detailed, accurate description of the Meselson–Stahl experiment.

A nice 7:30 min video describing the Meselson–Stahl experiment and its conclusions.

This film documents the discovery of the structure and replication of DNA including interviews with James Watson who, along with Crick, proposed the double helix model of DNA.

This activity is often used in conjunction with the short film The Double Helix. It introduces students to Meselson and Stahl experiment and helps them understand the concepts generated via those experimental results.

This collection of resources from HHMI Biointeractive addresses many of the major concepts surrounding DNA and its production, reading, and replication.

11.2 DNA Replication

Learning objectives.

By the end of this section, you will be able to:

  • Explain the meaning of semiconservative DNA replication
  • Explain why DNA replication is bidirectional and includes both a leading and lagging strand
  • Explain why Okazaki fragments are formed
  • Describe the process of DNA replication and the functions of the enzymes involved
  • Identify the differences between DNA replication in bacteria and eukaryotes
  • Explain the process of rolling circle replication

The elucidation of the structure of the double helix by James Watson and Francis Crick in 1953 provided a hint as to how DNA is copied during the process of replication . Separating the strands of the double helix would provide two templates for the synthesis of new complementary strands, but exactly how new DNA molecules were constructed was still unclear. In one model, semiconservative replication , the two strands of the double helix separate during DNA replication, and each strand serves as a template from which the new complementary strand is copied; after replication, each double-stranded DNA includes one parental or “old” strand and one “new” strand. There were two competing models also suggested: conservative and dispersive, which are shown in Figure 11.4 .

Matthew Meselson (1930–) and Franklin Stahl (1929–) devised an experiment in 1958 to test which of these models correctly represents DNA replication ( Figure 11.5 ). They grew E. coli for several generations in a medium containing a “heavy” isotope of nitrogen ( 15 N) that was incorporated into nitrogenous bases and, eventually, into the DNA. This labeled the parental DNA. The E. coli culture was then shifted into a medium containing 14 N and allowed to grow for one generation. The cells were harvested and the DNA was isolated. The DNA was separated by ultracentrifugation, during which the DNA formed bands according to its density. DNA grown in 15 N would be expected to form a band at a higher density position than that grown in 14 N. Meselson and Stahl noted that after one generation of growth in 14 N, the single band observed was intermediate in position in between DNA of cells grown exclusively in 15 N or 14 N. This suggested either a semiconservative or dispersive mode of replication. Some cells were allowed to grow for one more generation in 14 N and spun again. The DNA harvested from cells grown for two generations in 14 N formed two bands: one DNA band was at the intermediate position between 15 N and 14 N, and the other corresponded to the band of 14 N DNA. These results could only be explained if DNA replicates in a semiconservative manner. If DNA replication was dispersive, a single purple band positioned closer to the red 14 14 would have been observed, as more 14 was added in a dispersive manner to replace 15 . Therefore, the other two models were ruled out. As a result of this experiment, we now know that during DNA replication, each of the two strands that make up the double helix serves as a template from which new strands are copied. The new strand will be complementary to the parental or “old” strand. The resulting DNA molecules have the same sequence and are divided equally into the two daughter cells.

Check Your Understanding

  • What would have been the conclusion of Meselson and Stahl’s experiment if, after the first generation, they had found two bands of DNA?

DNA Replication in Bacteria

DNA replication has been well studied in bacteria primarily because of the small size of the genome and the mutants that are available. E. coli has 4.6 million base pairs (Mbp) in a single circular chromosome and all of it is replicated in approximately 42 minutes, starting from a single origin of replication and proceeding around the circle bidirectionally (i.e., in both directions). This means that approximately 1000 nucleotides are added per second. The process is quite rapid and occurs with few errors.

DNA replication uses a large number of proteins and enzymes ( Table 11.1 ). One of the key players is the enzyme DNA polymerase , also known as DNA pol. In bacteria, three main types of DNA polymerases are known: DNA pol I, DNA pol II, and DNA pol III. It is now known that DNA pol III is the enzyme required for DNA synthesis; DNA pol I and DNA pol II are primarily required for repair. DNA pol III adds deoxyribonucleotides each complementary to a nucleotide on the template strand, one by one to the 3’-OH group of the growing DNA chain. The addition of these nucleotides requires energy. This energy is present in the bonds of three phosphate groups attached to each nucleotide (a triphosphate nucleotide), similar to how energy is stored in the phosphate bonds of adenosine triphosphate (ATP) ( Figure 11.6 ). When the bond between the phosphates is broken and diphosphate is released, the energy released allows for the formation of a covalent phosphodiester bond by dehydration synthesis between the incoming nucleotide and the free 3’-OH group on the growing DNA strand.

The initiation of replication occurs at specific nucleotide sequence called the origin of replication , where various proteins bind to begin the replication process. E. coli has a single origin of replication (as do most prokaryotes), called oriC , on its one chromosome. The origin of replication is approximately 245 base pairs long and is rich in adenine-thymine (AT) sequences.

Some of the proteins that bind to the origin of replication are important in making single-stranded regions of DNA accessible for replication. Chromosomal DNA is typically wrapped around histones (in eukaryotes and archaea) or histone-like proteins (in bacteria), and is supercoiled , or extensively wrapped and twisted on itself. This packaging makes the information in the DNA molecule inaccessible. However, enzymes called topoisomerases change the shape and supercoiling of the chromosome. For bacterial DNA replication to begin, the supercoiled chromosome is relaxed by topoisomerase II , also called DNA gyrase . An enzyme called helicase then separates the DNA strands by breaking the hydrogen bonds between the nitrogenous base pairs. Recall that AT sequences have fewer hydrogen bonds and, hence, have weaker interactions than guanine-cytosine (GC) sequences. These enzymes require ATP hydrolysis. As the DNA opens up, Y-shaped structures called replication forks are formed. Two replication forks are formed at the origin of replication, allowing for bidirectional replication and formation of a structure that looks like a bubble when viewed with a transmission electron microscope; as a result, this structure is called a replication bubble . The DNA near each replication fork is coated with single-stranded binding proteins to prevent the single-stranded DNA from rewinding into a double helix.

Once single-stranded DNA is accessible at the origin of replication, DNA replication can begin. However, DNA pol III is able to add nucleotides only in the 5’ to 3’ direction (a new DNA strand can be only extended in this direction). This is because DNA polymerase requires a free 3’-OH group to which it can add nucleotides by forming a covalent phosphodiester bond between the 3’-OH end and the 5’ phosphate of the next nucleotide. This also means that it cannot add nucleotides if a free 3’-OH group is not available, which is the case for a single strand of DNA. The problem is solved with the help of an RNA sequence that provides the free 3’-OH end. Because this sequence allows the start of DNA synthesis, it is appropriately called the primer . The primer is five to 10 nucleotides long and complementary to the parental or template DNA. It is synthesized by RNA primase , which is an RNA polymerase . Unlike DNA polymerases, RNA polymerases do not need a free 3’-OH group to synthesize an RNA molecule. Now that the primer provides the free 3’-OH group, DNA polymerase III can now extend this RNA primer, adding DNA nucleotides one by one that are complementary to the template strand ( Figure 11.4 ).

During elongation in DNA replication , the addition of nucleotides occurs at its maximal rate of about 1000 nucleotides per second. DNA polymerase III can only extend in the 5’ to 3’ direction, which poses a problem at the replication fork. The DNA double helix is antiparallel; that is, one strand is oriented in the 5’ to 3’ direction and the other is oriented in the 3’ to 5’ direction (see Structure and Function of DNA ). During replication, one strand, which is complementary to the 3’ to 5’ parental DNA strand, is synthesized continuously toward the replication fork because polymerase can add nucleotides in this direction. This continuously synthesized strand is known as the leading strand . The other strand, complementary to the 5’ to 3’ parental DNA, grows away from the replication fork, so the polymerase must move back toward the replication fork to begin adding bases to a new primer, again in the direction away from the replication fork. It does so until it bumps into the previously synthesized strand and then it moves back again ( Figure 11.7 ). These steps produce small DNA sequence fragments known as Okazaki fragments , each separated by RNA primer. Okazaki fragments are named after the Japanese research team and married couple Reiji and Tsuneko Okazaki , who first discovered them in 1966. The strand with the Okazaki fragments is known as the lagging strand , and its synthesis is said to be discontinuous.

The leading strand can be extended from one primer alone, whereas the lagging strand needs a new primer for each of the short Okazaki fragments. The overall direction of the lagging strand will be 3’ to 5’, and that of the leading strand 5’ to 3’. A protein called the sliding clamp holds the DNA polymerase in place as it continues to add nucleotides. The sliding clamp is a ring-shaped protein that binds to the DNA and holds the polymerase in place. Beyond its role in initiation, topoisomerase also prevents the overwinding of the DNA double helix ahead of the replication fork as the DNA is opening up; it does so by causing temporary nicks in the DNA helix and then resealing it. As synthesis proceeds, the RNA primers are replaced by DNA. The primers are removed by the exonuclease activity of DNA polymerase I, and the gaps are filled in. The nicks that remain between the newly synthesized DNA (that replaced the RNA primer) and the previously synthesized DNA are sealed by the enzyme DNA ligase that catalyzes the formation of covalent phosphodiester linkage between the 3’-OH end of one DNA fragment and the 5’ phosphate end of the other fragment, stabilizing the sugar-phosphate backbone of the DNA molecule.

Termination

Once the complete chromosome has been replicated, termination of DNA replication must occur. Although much is known about initiation of replication, less is known about the termination process. Following replication, the resulting complete circular genomes of prokaryotes are concatenated, meaning that the circular DNA chromosomes are interlocked and must be separated from each other. This is accomplished through the activity of bacterial topoisomerase IV, which introduces double-stranded breaks into DNA molecules, allowing them to separate from each other; the enzyme then reseals the circular chromosomes. The resolution of concatemers is an issue unique to prokaryotic DNA replication because of their circular chromosomes. Because both bacterial DNA gyrase and topoisomerase IV are distinct from their eukaryotic counterparts, these enzymes serve as targets for a class of antimicrobial drugs called quinolones .

  • Which enzyme breaks the hydrogen bonds holding the two strands of DNA together so that replication can occur?
  • Is it the lagging strand or the leading strand that is synthesized in the direction toward the opening of the replication fork?
  • Which enzyme is responsible for removing the RNA primers in newly replicated bacterial DNA?

DNA Replication in Eukaryotes

Eukaryotic genomes are much more complex and larger than prokaryotic genomes and are typically composed of multiple linear chromosomes ( Table 11.2 ). The human genome , for example, has 3 billion base pairs per haploid set of chromosomes, and 6 billion base pairs are inserted during replication. There are multiple origins of replication on each eukaryotic chromosome ( Figure 11.8 ); the human genome has 30,000 to 50,000 origins of replication. The rate of replication is approximately 100 nucleotides per second—10 times slower than prokaryotic replication.

The essential steps of replication in eukaryotes are the same as in prokaryotes. Before replication can start, the DNA has to be made available as a template. Eukaryotic DNA is highly supercoiled and packaged, which is facilitated by many proteins, including histone s (see Structure and Function of Cellular Genomes ). At the origin of replication , a prereplication complex composed of several proteins, including helicase , forms and recruits other enzymes involved in the initiation of replication, including topoisomerase to relax supercoiling, single-stranded binding protein, RNA primase , and DNA polymerase . Following initiation of replication, in a process similar to that found in prokaryotes, elongation is facilitated by eukaryotic DNA polymerases. The leading strand is continuously synthesized by the eukaryotic polymerase enzyme pol δ, while the lagging strand is synthesized by pol ε. A sliding clamp protein holds the DNA polymerase in place so that it does not fall off the DNA. The enzyme ribonuclease H ( RNase H ), instead of a DNA polymerase as in bacteria, removes the RNA primer, which is then replaced with DNA nucleotides. The gaps that remain are sealed by DNA ligase .

Because eukaryotic chromosomes are linear, one might expect that their replication would be more straightforward. As in prokaryotes, the eukaryotic DNA polymerase can add nucleotides only in the 5’ to 3’ direction. In the leading strand, synthesis continues until it reaches either the end of the chromosome or another replication fork progressing in the opposite direction. On the lagging strand, DNA is synthesized in short stretches, each of which is initiated by a separate primer. When the replication fork reaches the end of the linear chromosome, there is no place to make a primer for the DNA fragment to be copied at the end of the chromosome. These ends thus remain unpaired and, over time, they may get progressively shorter as cells continue to divide.

The ends of the linear chromosomes are known as telomere s and consist of noncoding repetitive sequences. The telomeres protect coding sequences from being lost as cells continue to divide. In humans, a six base-pair sequence, TTAGGG, is repeated 100 to 1000 times to form the telomere. The discovery of the enzyme telomerase ( Figure 11.9 ) clarified our understanding of how chromosome ends are maintained. Telomerase contains a catalytic part and a built-in RNA template. It attaches to the end of the chromosome, and complementary bases to the RNA template are added on the 3’ end of the DNA strand. Once the 3’ end of the lagging strand template is sufficiently elongated, DNA polymerase can add the nucleotides complementary to the ends of the chromosomes. In this way, the ends of the chromosomes are replicated. In humans, telomerase is typically active in germ cells and adult stem cells; it is not active in adult somatic cells and may be associated with the aging of these cells. Eukaryotic microbes including fungi and protozoans also produce telomerase to maintain chromosomal integrity. For her discovery of telomerase and its action, Elizabeth Blackburn (1948–) received the Nobel Prize for Medicine or Physiology in 2009.

Link to Learning

This animation compares the process of prokaryotic and eukaryotic DNA replication.

  • How does the origin of replication differ between eukaryotes and prokaryotes?
  • What polymerase enzymes are responsible for DNA synthesis during eukaryotic replication?
  • What is found at the ends of the chromosomes in eukaryotes and why?

DNA Replication of Extrachromosomal Elements: Plasmids and Viruses

To copy their nucleic acids, plasmids and viruses frequently use variations on the pattern of DNA replication described for prokaryote genomes. For more information on the wide range of viral replication strategies, see The Viral Life Cycle .

Rolling Circle Replication

Whereas many bacterial plasmids (see Unique Characteristics of Prokaryotic Cells ) replicate by a process similar to that used to copy the bacterial chromosome, other plasmids, several bacteriophages , and some viruses of eukaryotes use rolling circle replication ( Figure 11.10 ). The circular nature of plasmids and the circularization of some viral genomes on infection make this possible. Rolling circle replication begins with the enzymatic nicking of one strand of the double-stranded circular molecule at the double-stranded origin (dso) site . In bacteria, DNA polymerase III binds to the 3’-OH group of the nicked strand and begins to unidirectionally replicate the DNA using the un-nicked strand as a template, displacing the nicked strand as it does so. Completion of DNA replication at the site of the original nick results in full displacement of the nicked strand, which may then recircularize into a single-stranded DNA molecule. RNA primase then synthesizes a primer to initiate DNA replication at the single-stranded origin (sso) site of the single-stranded DNA (ssDNA) molecule, resulting in a double-stranded DNA (dsDNA) molecule identical to the other circular DNA molecule.

  • Is there a lagging strand in rolling circle replication? Why or why not?

As an Amazon Associate we earn from qualifying purchases.

This book may not be used in the training of large language models or otherwise be ingested into large language models or generative AI offerings without OpenStax's permission.

Want to cite, share, or modify this book? This book uses the Creative Commons Attribution License and you must attribute OpenStax.

Access for free at https://openstax.org/books/microbiology/pages/1-introduction
  • Authors: Nina Parker, Mark Schneegurt, Anh-Hue Thi Tu, Philip Lister, Brian M. Forster
  • Publisher/website: OpenStax
  • Book title: Microbiology
  • Publication date: Nov 1, 2016
  • Location: Houston, Texas
  • Book URL: https://openstax.org/books/microbiology/pages/1-introduction
  • Section URL: https://openstax.org/books/microbiology/pages/11-2-dna-replication

© Jan 10, 2024 OpenStax. Textbook content produced by OpenStax is licensed under a Creative Commons Attribution License . The OpenStax name, OpenStax logo, OpenStax book covers, OpenStax CNX name, and OpenStax CNX logo are not subject to the Creative Commons license and may not be reproduced without the prior and express written consent of Rice University.

Library homepage

  • school Campus Bookshelves
  • menu_book Bookshelves
  • perm_media Learning Objects
  • login Login
  • how_to_reg Request Instructor Account
  • hub Instructor Commons
  • Download Page (PDF)
  • Download Full Book (PDF)
  • Periodic Table
  • Physics Constants
  • Scientific Calculator
  • Reference & Cite
  • Tools expand_more
  • Readability

selected template will load here

This action is not available.

Biology LibreTexts

9.2: DNA Replication

  • Last updated
  • Save as PDF
  • Page ID 7023

When a cell divides, it is important that each daughter cell receives an identical copy of the DNA. This is accomplished by the process of DNA replication. The replication of DNA occurs during the synthesis phase, or S phase, of the cell cycle, before the cell enters mitosis or meiosis.

The elucidation of the structure of the double helix provided a hint as to how DNA is copied. Recall that adenine nucleotides pair with thymine nucleotides, and cytosine with guanine. This means that the two strands are complementary to each other. For example, a strand of DNA with a nucleotide sequence of AGTCATGA will have a complementary strand with the sequence TCAGTACT (Figure \(\PageIndex{1}\)).

Figure shows the ladder-like structure of DNA, with complementary bases making up the rungs of the ladder.

Because of the complementarity of the two strands, having one strand means that it is possible to recreate the other strand. This model for replication suggests that the two strands of the double helix separate during replication, and each strand serves as a template from which the new complementary strand is copied (Figure \(\PageIndex{2}\)).

Illustration shows the semiconservative model of DNA synthesis. In the semi-conservative model, each newly synthesized strand pairs with a parent strand.

During DNA replication, each of the two strands that make up the double helix serves as a template from which new strands are copied. The new strand will be complementary to the parental or “old” strand. Each new double strand consists of one parental strand and one new daughter strand. This is known as semiconservative replication. When two DNA copies are formed, they have an identical sequence of nucleotide bases and are divided equally into two daughter cells.

DNA Replication in Eukaryotes

Because eukaryotic genomes are very complex, DNA replication is a very complicated process that involves several enzymes and other proteins. It occurs in three main stages: initiation, elongation, and termination.

Recall that eukaryotic DNA is bound to proteins known as histones to form structures called nucleosomes. During initiation, the DNA is made accessible to the proteins and enzymes involved in the replication process. How does the replication machinery know where on the DNA double helix to begin? It turns out that there are specific nucleotide sequences called origins of replication at which replication begins. Certain proteins bind to the origin of replication while an enzyme called helicase unwinds and opens up the DNA helix. As the DNA opens up, Y-shaped structures called replication forks are formed (Figure \(\PageIndex{3}\)). Two replication forks are formed at the origin of replication, and these get extended in both directions as replication proceeds. There are multiple origins of replication on the eukaryotic chromosome, such that replication can occur simultaneously from several places in the genome.

During elongation, an enzyme called DNA polymerase adds DNA nucleotides to the 3' end of the template. Because DNA polymerase can only add new nucleotides at the end of a backbone, a primer sequence, which provides this starting point, is added with complementary RNA nucleotides. This primer is removed later, and the nucleotides are replaced with DNA nucleotides. One strand, which is complementary to the parental DNA strand, is synthesized continuously toward the replication fork so the polymerase can add nucleotides in this direction. This continuously synthesized strand is known as the leading strand. Because DNA polymerase can only synthesize DNA in a 5' to 3' direction, the other new strand is put together in short pieces called Okazaki fragments. The Okazaki fragments each require a primer made of RNA to start the synthesis. The strand with the Okazaki fragments is known as the lagging strand. As synthesis proceeds, an enzyme removes the RNA primer, which is then replaced with DNA nucleotides, and the gaps between fragments are sealed by an enzyme called DNA ligase.

The process of DNA replication can be summarized as follows:

  • DNA unwinds at the origin of replication.
  • New bases are added to the complementary parental strands. One new strand is made continuously, while the other strand is made in pieces.
  • Primers are removed, new DNA nucleotides are put in place of the primers and the backbone is sealed by DNA ligase.

ART CONNECTION

Illustration shows a replication bubble. Helicase unwinds the helix. An RNA primer starts the synthesis, and DNA polymerase extends the DNA strand from the RNA primer. DNA synthesis occurs only in the 5' to 3' direction. On the leading strand, DNA synthesis occurs continuously. On the lagging strand, DNA synthesis restarts many times as the helix unwinds, resulting in many short fragments called Okazaki fragments.

You isolate a cell strain in which the joining together of Okazaki fragments is impaired and suspect that a mutation has occurred in an enzyme found at the replication fork. Which enzyme is most likely to be mutated?

Telomere Replication

Because eukaryotic chromosomes are linear, DNA replication comes to the end of a line in eukaryotic chromosomes. As you have learned, the DNA polymerase enzyme can add nucleotides in only one direction. In the leading strand, synthesis continues until the end of the chromosome is reached; however, on the lagging strand there is no place for a primer to be made for the DNA fragment to be copied at the end of the chromosome. This presents a problem for the cell because the ends remain unpaired, and over time these ends get progressively shorter as cells continue to divide. The ends of the linear chromosomes are known as telomeres, which have repetitive sequences that do not code for a particular gene. As a consequence, it is telomeres that are shortened with each round of DNA replication instead of genes. For example, in humans, a six base-pair sequence, TTAGGG, is repeated 100 to 1000 times. The discovery of the enzyme telomerase (Figure \(\PageIndex{4}\)) helped in the understanding of how chromosome ends are maintained. The telomerase attaches to the end of the chromosome, and complementary bases to the RNA template are added on the end of the DNA strand. Once the lagging strand template is sufficiently elongated, DNA polymerase can now add nucleotides that are complementary to the ends of the chromosomes. Thus, the ends of the chromosomes are replicated.

Telomerase has an associated RNA that complements the 5' overhang at the end of the chromosome. The RNA template is used to synthesize the complementary strand. Telomerase then shifts, and the process is repeated. Next, primase and DNA polymerase synthesize the rest of the complementary strand.

Telomerase is typically found to be active in germ cells, adult stem cells, and some cancer cells. For her discovery of telomerase and its action, Elizabeth Blackburn (Figure \(\PageIndex{5}\)) received the Nobel Prize for Medicine and Physiology in 2009.

Photo shows Elizabeth Blackburn.

Telomerase is not active in adult somatic cells. Adult somatic cells that undergo cell division continue to have their telomeres shortened. This essentially means that telomere shortening is associated with aging. In 2010, scientists found that telomerase can reverse some age-related conditions in mice, and this may have potential in regenerative medicine. 1 Telomerase-deficient mice were used in these studies; these mice have tissue atrophy, stem-cell depletion, organ system failure, and impaired tissue injury responses. Telomerase reactivation in these mice caused extension of telomeres, reduced DNA damage, reversed neurodegeneration, and improved functioning of the testes, spleen, and intestines. Thus, telomere reactivation may have potential for treating age-related diseases in humans.

DNA Replication in Prokaryotes

Recall that the prokaryotic chromosome is a circular molecule with a less extensive coiling structure than eukaryotic chromosomes. The eukaryotic chromosome is linear and highly coiled around proteins. While there are many similarities in the DNA replication process, these structural differences necessitate some differences in the DNA replication process in these two life forms.

DNA replication has been extremely well-studied in prokaryotes, primarily because of the small size of the genome and large number of variants available. Escherichia coli has 4.6 million base pairs in a single circular chromosome, and all of it gets replicated in approximately 42 minutes, starting from a single origin of replication and proceeding around the chromosome in both directions. This means that approximately 1000 nucleotides are added per second. The process is much more rapid than in eukaryotes. Table \(\PageIndex{1}\) summarizes the differences between prokaryotic and eukaryotic replications.

CONCEPT IN ACTION

Click through a tutorial on DNA replication.

DNA polymerase can make mistakes while adding nucleotides. It edits the DNA by proofreading every newly added base. Incorrect bases are removed and replaced by the correct base, and then polymerization continues (Figure \(\PageIndex{6}\) a ). Most mistakes are corrected during replication, although when this does not happen, the mismatch repair mechanism is employed. Mismatch repair enzymes recognize the wrongly incorporated base and excise it from the DNA, replacing it with the correct base (Figure \(\PageIndex{6}\) b ). In yet another type of repair, nucleotide excision repair, the DNA double strand is unwound and separated, the incorrect bases are removed along with a few bases on the 5' and 3' end, and these are replaced by copying the template with the help of DNA polymerase (Figure \(\PageIndex{6}\) c ). Nucleotide excision repair is particularly important in correcting thymine dimers, which are primarily caused by ultraviolet light. In a thymine dimer, two thymine nucleotides adjacent to each other on one strand are covalently bonded to each other rather than their complementary bases. If the dimer is not removed and repaired it will lead to a mutation. Individuals with flaws in their nucleotide excision repair genes show extreme sensitivity to sunlight and develop skin cancers early in life.

 Part a shows DNA polymerase replicating a strand of DNA. The enzyme has accidentally inserted G opposite A, resulting in a bulge. The enzyme backs up to fix the error. In part b, the top illustration shows a replicated DNA strand with a G–T base mismatch. The bottom illustration shows the repaired DNA, which has the correct G–C base pairing. Part c shows  a DNA strand in which a thymine dimer has formed. An excision repair enzyme cuts out the section of DNA that contains the dimer so that it can be replaced with a normal base pair.

Most mistakes are corrected; if they are not, they may result in a mutation—defined as a permanent change in the DNA sequence. Mutations in repair genes may lead to serious consequences like cancer.

DNA replicates by a semi-conservative method in which each of the two parental DNA strands act as a template for new DNA to be synthesized. After replication, each DNA has one parental or “old” strand, and one daughter or “new” strand.

Replication in eukaryotes starts at multiple origins of replication, while replication in prokaryotes starts from a single origin of replication. The DNA is opened with enzymes, resulting in the formation of the replication fork. Primase synthesizes an RNA primer to initiate synthesis by DNA polymerase, which can add nucleotides in only one direction. One strand is synthesized continuously in the direction of the replication fork; this is called the leading strand. The other strand is synthesized in a direction away from the replication fork, in short stretches of DNA known as Okazaki fragments. This strand is known as the lagging strand. Once replication is completed, the RNA primers are replaced by DNA nucleotides and the DNA is sealed with DNA ligase.

The ends of eukaryotic chromosomes pose a problem, as polymerase is unable to extend them without a primer. Telomerase, an enzyme with an inbuilt RNA template, extends the ends by copying the RNA template and extending one end of the chromosome. DNA polymerase can then extend the DNA using the primer. In this way, the ends of the chromosomes are protected. Cells have mechanisms for repairing DNA when it becomes damaged or errors are made in replication. These mechanisms include mismatch repair to replace nucleotides that are paired with a non-complementary base and nucleotide excision repair, which removes bases that are damaged such as thymine dimers.

Art Connections

Figure \(\PageIndex{3}\): You isolate a cell strain in which the joining together of Okazaki fragments is impaired and suspect that a mutation has occurred in an enzyme found at the replication fork. Which enzyme is most likely to be mutated?

Ligase, as this enzyme joins together Okazaki fragments.

  • 1 Mariella Jaskelioff, et al., “Telomerase reactivation reverses tissue degeneration in aged telomerase-deficient mice,” Nature , 469 (2011):102–7.

Contributors and Attributions

Samantha Fowler (Clayton State University), Rebecca Roush (Sandhills Community College), James Wise (Hampton University). Original content by OpenStax (CC BY 4.0; Access for free at https://cnx.org/contents/b3c1e1d2-83...4-e119a8aafbdd ).

This page has been archived and is no longer updated

Semi-Conservative DNA Replication: Meselson and Stahl

hypothesis on dna replication

This structure has novel features which are of considerable biological interest . . . It has not escaped our notice that the specific pairing we have postulated immediately suggests a possible copying mechanism for the genetic material . —Watson & Crick (1953)

Perhaps the most significant aspect of Watson and Crick's discovery of DNA structure was not that it provided scientists with a three-dimensional model of this molecule , but rather that this structure seemed to reveal the way in which DNA was replicated. As noted in their 1953 paper, Watson and Crick strongly suspected that the specific base pairings within the DNA double helix existed in order to ensure a controlled system of DNA replication . However, it took several years of subsequent study, including a classic 1958 experiment by American geneticists Matthew Meselson and Franklin Stahl, before the exact relationship between DNA structure and replication was understood.

Three Proposed Models for DNA Replication

Replication is the process by which a cell copies its DNA prior to division. In humans, for example, each parent cell must copy its entire six billion base pairs of DNA before undergoing mitosis . The molecular details of DNA replication are described elsewhere, and they were not known until some time after Watson and Crick's discovery. In fact, before such details could be determined, scientists were faced with a more fundamental research concern. Specifically, they wanted to know the overall nature of the process by which DNA replication occurs.

Defining the Models

Semiconservative replication was not the only model of DNA replication proposed during the mid-1950s, however. In fact, two other prominent hypotheses were put also forth: conservative replication and dispersive replication. According to the conservative replication model, the entire original DNA double helix serves as a template for a new double helix, such that each round of cell division produces one daughter cell with a completely new DNA double helix and another daughter cell with a completely intact old (or original) DNA double helix. On the other hand, in the dispersive replication model, the original DNA double helix breaks apart into fragments, and each fragment then serves as a template for a new DNA fragment. As a result, every cell division produces two cells with varying amounts of old and new DNA (Figure 1).

Making Predictions Based on the Models

Meselson and stahl’s elegant experiment.

The duo thus began their experiment by choosing two isotopes of nitrogen—the common and lighter 14 N, and the rare and heavier 15 N (so-called "heavy" nitrogen)—as their labels and a technique known as cesium chloride (CsCl) equilibrium density gradient centrifugation as their sedimentation method. Meselson and Stahl opted for nitrogen because it is an essential chemical component of DNA; therefore, every time a cell divides and its DNA replicates, it incorporates new N atoms into the DNA of either one or both of its two daughter cells, depending on which model was correct. "If several different density species of DNA are present," they predicted, "each will form a band at the position where the density of the CsCl solution is equal to the buoyant density of that species. In this way, DNA labeled with heavy nitrogen ( 15 N) may be resolved from unlabeled DNA" (Meselson & Stahl, 1958).

The scientists then continued their experiment by growing a culture of E. coli bacteria in a medium that had the heavier 15 N (in the form of 15 N-labeled ammonium chloride) as its only source of nitrogen. In fact, they did this for 14 bacterial generations, which was long enough to create a population of bacterial cells that contained only the heavier isotope (all the original 14 N-containing cells had died by then). Next, they changed the medium to one containing only 14 N-labeled ammonium salts as the sole nitrogen source. So, from that point onward, every new strand of DNA would be built with 14 N rather than 15 N.

Just prior to the addition of 14 N and periodically thereafter, as the bacterial cells grew and replicated, Meselson and Stahl sampled DNA for use in equilibrium density gradient centrifugation to determine how much 15 N (from the original or old DNA) versus 14 N (from the new DNA) was present. For the centrifugation procedure, they mixed the DNA samples with a solution of cesium chloride and then centrifuged the samples for enough time to allow the heavier 15 N and lighter 14 N DNA to migrate to different positions in the centrifuge tube.

By way of centrifugation, the scientists found that DNA composed entirely of 15 N-labeled DNA (i.e., DNA collected just prior to changing the culture from one containing only 15 N to one containing only 14 N) formed a single distinct band, because both of its strands were made entirely in the "heavy" nitrogen medium. Following a single round of replication, the DNA again formed a single distinct band, but the band was located in a different position along the centrifugation gradient. Specifically, it was found midway between where all the 15 N and all the 14 N DNA would have migrated—in other words, halfway between "heavy" and "light" (Figure 2). Based on these findings, the scientists were immediately able to exclude the conservative model of replication as a possibility. After all, if DNA replicated conservatively, there should have been two distinct bands after a single round of replication; half of the new DNA would have migrated to the same position as it did before the culture was transferred to the 14 N-containing medium (i.e., to the "heavy" position), and only the other half would have migrated to the new position (i.e., to the "light" position). That left the scientists with only two options: either DNA replicated semiconservatively, as Watson and Crick had predicted, or it replicated dispersively.

Straight or Circular?

Following publication of Meselson and Stahl's results, many scientists confirmed that semiconservative replication was the rule, not just in E. coli , but in every other species studied as well. To date, no one has found any evidence for either conservative or dispersive DNA replication. Scientists have found, however, that semiconservative replication can occur in different ways—for example, it may proceed in either a circular or a linear fashion, depending on chromosome shape.

In fact, in the early 1960s, English molecular biologist John Cairns performed another remarkably elegant experiment to demonstrate that E. coli and other bacteria with circular chromosomes undergo what he termed " theta replication ," because the structure generated resembles the Greek letter theta (Θ). Specifically, Cairns grew E. coli bacteria in the presence of radioactive nucleotides such that, after replication, each new DNA molecule had one radioactive (hot) strand and one nonradioactive strand. He then isolated the newly replicated DNA and used it to produce an electron micrograph image of the Θ-shaped replication process (Figure 3; Cairns, 1961).

References and Recommended Reading

Cairns, J. The bacterial chromosome and its manner of replication as seen by autoradiography. Journal of Molecular Biology 6 , 208–213 (1961)

Meselson, M., & Stahl, F. The replication of DNA in Escherichia coli . Proceedings of the National Academy of Sciences 44 , 671–682 (1958)

Watson, J. D., & Crick, F. H. C. A structure for deoxyribose nucleic acid. Nature 171 , 737–738 (1953) ( link to article ).

  • Add Content to Group

Article History

Flag inappropriate.

Google Plus+

StumbleUpon

Email your Friend

hypothesis on dna replication

  •  |  Lead Editor:  Bob Moss

Topic Rooms

Within this Subject (34)

  • Applications in Biotechnology (4)
  • Discovery of Genetic Material (4)
  • DNA Replication (6)
  • Gene Copies (5)
  • Jumping Genes (4)
  • RNA (7)
  • Transcription & Translation (4)

Other Topic Rooms

  • Gene Inheritance and Transmission
  • Gene Expression and Regulation
  • Nucleic Acid Structure and Function
  • Chromosomes and Cytogenetics
  • Evolutionary Genetics
  • Population and Quantitative Genetics
  • Genes and Disease
  • Genetics and Society
  • Cell Origins and Metabolism
  • Proteins and Gene Expression
  • Subcellular Compartments
  • Cell Communication
  • Cell Cycle and Cell Division

ScholarCast

© 2014 Nature Education

  • Press Room |
  • Terms of Use |
  • Privacy Notice |

Send

Visual Browse

  • Illustrations
  • Problem Sets
  • Calculators

You are here

Meselson–stahl experiment.

In their second paper on the structure of DNA * , Watson and Crick (pdf) described how DNA's structure suggests a pattern for replication:

"…prior to duplication the hydrogen bonds are broken, and the two chains unwind and separate. Each chain then acts as a template for the formation onto itself of a new companion chain, so that eventually we shall have two pairs of chains, where we only had one before." - Watson and Crick, 1953

This is called semiconservative replication .

Today we know that this is the pattern used by living cells, but the experimental evidence in support of semiconservative replication was not published until 1958 . In the 5 years between Watson and Crick's suggestion and the definitive experiment, semiconservative replication was controversial and other patterns were considered.

Three hypothesized patterns were proposed:

  • Semiconservative - The original double strand of DNA separates and each strand acts as a template for the synthesis of a complimentary strand.
  • Conservative replication - the original double strand of DNA remains intact and is used as a template to create a new double stranded molecule.
  • Dispersive replication - similar to conservative replication in that the original double strand is used as a template without being separated, but prior to cell division, the strands recombine such that each daughter cell gets a mix of new and old DNA. With each round of replication, the original DNA gets cut up and dispersed evenly between each copy.

The methods Meselson and Stahl developed allowed them to distinguish existing DNA from newly synthesized DNA and to track new and old DNA over several rounds of replication.

They accomplished this by labeling cells with different stable isotopes of nitrogen. First, a culture of bacterial cell were grown for several generations in a media containing only 15 N ( a stable, heavy isotope of Nitrogen). After this period * of growth, all of the DNA in the cells contained 15 N. These cells were then rinsed and put into a media containing only the more common, lighter isotope of nitrogen ( 14 N). As the cells grew and divided in this fresh media, all newly synthesized DNA would contain only the lighter nitrogen isotope, while DNA from the original cells would still contain 15 N. In this illustration above, 15 N labeled DNA is shown in orange and 14 N labeled in green.

The 15 N and 14 N labeled DNA was then tracked using high speed centrifugation and a density * gradient created with cesium chloride (CsCl).

During centrifugation in a CsCl gradient, DNA accumulates in bands along the gradient based on its density. Since 15 N is more dense than 14 N, 15 N enriched DNA accumulates lower down in the centrifuge tube than the 14 N DNA. DNA containing a mixture of 15 N and 14 N ends up in an intermediate position between the two extremes.

By spinning DNA extracted at different times during the experiment, Meselson and Stahl were able to see how new and old DNA interacted during each round of replication.

The beauty of this experiment was that it allowed them to distinguish between the three different hypothesized replication patterns. The key result occurs at the second generation when all three proposed replication patterns give different results in the CsCl gradient.

That Meselson and Stahl's experiment showed the pattern predicted by the semiconservative hypothesis provided the definitive experimental evidence in support of the process proposed by Watson and Crick.

Related Content

  • Complementary Nucleotide Bases
  • DNA Polymerase
  • Semiconservative Replication
  • Nucleotides in DNA practice problems
  • Complementary nucleotide bases concept questions
  • DNA Polymerase concept questions
  • DNA replication

Subject tag:

  • Printer-friendly version

Search form

Latest illustrations.

  • Ocean Gyres and Geostrophic Flow
  • Langmuir Circulation
  • Test sensitivity - specificity calculator
  • How earthquakes show us the inside of the Earth
  • Surface currents, the Ekman spiral, and Ekman transport

© Andrew Staroscik 2011-2022

IMAGES

  1. Hypotheses on DNA Replication Animation Model

    hypothesis on dna replication

  2. PPT

    hypothesis on dna replication

  3. DNA replication

    hypothesis on dna replication

  4. Chapter: DNA Synthesis

    hypothesis on dna replication

  5. ? Replication of DNA

    hypothesis on dna replication

  6. DNA replication steps and rules, DNA polymerase enzymes and RNA primer

    hypothesis on dna replication

VIDEO

  1. Mathematical problems on DNA Replication

  2. Whether DNA existed first or RNA| Know your facts with AgriAddict

  3. Wobble hypothesis explained in bangla || Biochemistry

  4. PART 7: CELL CYCLE & PROTEIN SYNTHESIS_HYPOTHESIS RELATED TO DNA REPLICATION

  5. Translation (Protein synthesis) explained in bangla || Biochemistry

  6. Replication of DNA is? #neet #biology #mcq

COMMENTS

  1. Mode of DNA replication: Meselson-Stahl experiment - Khan Academy

    Dispersive replication. In the dispersive model, DNA replication results in two DNA molecules that are mixtures, or “hybrids,” of parental and daughter DNA. In this model, each individual strand is a patchwork of original and new DNA. Most biologists at the time would likely have put their money on the semi-conservative model.

  2. How DNA Replicates | Matthew Meselson | Franklin W. Stahl

    Max Delbruck, the Caltech physicist and biologist who had proposed the dispersive model, was elated by the results, even though Meselson and Stahl disproved his replication hypothesis, and urged the young scientists to write up their results for publication and announce the important result to the world (1958).

  3. Meselson–Stahl experiment - Wikipedia

    The Meselson–Stahl experiment is an experiment by Matthew Meselson and Franklin Stahl in 1958 which supported Watson and Crick 's hypothesis that DNA replication was semiconservative. In semiconservative replication, when the double-stranded DNA helix is replicated, each of the two new double-stranded DNA helices consisted of one strand from ...

  4. 11.2 DNA Replication - Microbiology | OpenStax

    Figure 11.7 At the origin of replication, topoisomerase II relaxes the supercoiled chromosome. Two replication forks are formed by the opening of the double-stranded DNA at the origin, and helicase separates the DNA strands, which are coated by single-stranded binding proteins to keep the strands separated.

  5. 7.3A: Basics of DNA Replication - Biology LibreTexts

    Basics of DNA Replication. Watson and Crick’s discovery that DNA was a two-stranded double helix provided a hint as to how DNA is replicated. During cell division, each DNA molecule has to be perfectly copied to ensure identical DNA molecules to move to each of the two daughter cells. The double-stranded structure of DNA suggested that the ...

  6. 9.2: DNA Replication - Biology LibreTexts

    9.2: DNA Replication. When a cell divides, it is important that each daughter cell receives an identical copy of the DNA. This is accomplished by the process of DNA replication. The replication of DNA occurs during the synthesis phase, or S phase, of the cell cycle, before the cell enters mitosis or meiosis. The elucidation of the structure of ...

  7. Semi-Conservative DNA Replication | Learn Science at Scitable

    Replication is the process by which a cell copies its DNA prior to division. In humans, for example, each parent cell must copy its entire six billion base pairs of DNA before undergoing mitosis ...

  8. Meselson–Stahl Experiment | Science Primer

    Support for the dispersive hypothesis remained strong until proof of semiconservative replication was provided by Meselson and stahl's 1958 paper (pdf). The methods Meselson and Stahl developed allowed them to distinguish existing DNA from newly synthesized DNA and to track new and old DNA over several rounds of replication.