Unraveling the Mystery: The Amazing Proof of the Poincaré Conjecture

Discover the incredible proof that solved one of math's greatest mysteries. Learn about the impact of the Poincaré Conjecture on modern geometry and topology.

Unraveling the Mystery: The Amazing Proof of the Poincaré Conjecture
Observers view inside a three-torus.

Hey math enthusiasts, have you ever wondered what it takes to crack a seemingly unsolvable mathematical puzzle? Hold onto your hats as I'm about to take you on an exhilarating ride through the astonishing proof of the Poincare Conjecture. This mathematical enigma has perplexed the brightest minds for over a century, but thanks to the genius of one man building upon the brilliant work of others alike, the world finally has the answer.

How the Poincaré Conjecture Came to Be

In the infinite expanse of the world of mathematics, there exists many luminous stars that have contributed greatly to the field. Among them is the great mathematician Jules Henri Poincaré. Born in 1854 in the picturesque town of Nancy, France, Poincare was destined to become one of the most prolific mathematicians and theoretical physicists of his time.

Figure 1: Jules Henri Poincaré in a photograph published in 1913. 

 In the landscape of Poincaré's many accomplishments, one observes a rich garden of interwoven branches, each bearing the fruits of his intellect. The seeds of his passion for mathematics were sown early, as he demonstrated a prodigious aptitude for the subject as a child. Nurturing this gift, Poincare went on to study at the prestigious École Polytechnique and École des Mines, where he first made his mark in the mathematical realm.

Figure 2: Front view of École Polytechnique

 After completing his studies at the École des Mines, Poincaré began his academic career, eventually obtaining a position at the University of Paris, where he would remain for the rest of his life.

Figure 3: La Sorbonne (Paris 1) which was part of the original University of Paris

 Perhaps one of Poincaré earliest accomplishments was his work on the three-body problem. In his research on this problem, Poincaré successfully established the qualitative theory of differential equations and discovered the existence of chaotic behavior in the solar system. This work, which earned him the prestigious King Oscar II Prize in 1889, led to the development of the field of dynamical systems and laid the groundwork for modern chaos theory.

 Poincaré was also heavily engaged in philosophical questions related to the foundations of mathematics and science. He was an early proponent of the idea that mathematics is a human creation, and that its effectiveness in describing the physical world is, to some extent, a mystery. Poincaré's philosophical work, particularly his writings on the nature of mathematics, have had a lasting impact on the philosophy of mathematics and the philosophy of science.

 Poincare's journey through the mathematical cosmos didn't stop there, as he was also involved with the discovery of non-Euclidean geometry. Here, he delved deep into the realm of hyperbolic space, unveiling the beauty of unfamiliar and mind-bending geometry. Through his work on this subject, he demonstrated the consistency of certain properties in non-Euclidean geometry, thereby revolutionizing our understanding of the very fabric of space which was previously thought to only exist within Euclidean space.

 Furthermore, Poincaré made huge contributions that greatly progressed the field of theoretical physics. In the nascent years of the 20th century, when the world of physics was undergoing a seismic shift, Poincare's work on the Lorentz transformations and the concept of space-time would prove instrumental in shaping the foundations and laying the framework for Einstein's theory of special relativity.

 Finally, we arrive at the grand crescendo of his work—the Poincaré Conjecture. Proposed in 1904, the conjecture is a statement about three-dimensional spaces, specifically those of a topological nature. In his conjecture, Poincaré posited that every simply connected, closed three-dimensional manifold is, in fact, homeomorphic to the three-dimensional sphere. In simpler terms, if you can smoothly deform a shape without tearing or gluing it, and it has no holes, then it must be equivalent to a sphere. This conjecture would mark the grand finale of the symphony of Poincaré's genius before his passing in 1912, tantalizing even the greatest mathematicians for over a century leaving behind Poincaré's influence on topology as an interminable one.

 The depth of Poincaré's understanding of mathematics and physics can be seen in his keen intuition and ability to synthesize ideas from various branches of knowledge allowing him to make groundbreaking strides in various different fields. This was why he was often described as the last polymath, and in mathematics as "The Last Universalist", since he excelled in all fields of the discipline as it existed during his lifetime.

Understanding the Poincaré Conjecture

To delve into the Poincare Conjecture, we must first acquaint ourselves with some essential concepts in topology. Topology is a fascinating branch of mathematics that deals with the study of shapes and spaces. It's kind of like geometry, but instead of focusing on angles, distances, and the like, topology zeros in on more abstract properties of shapes – things like connectivity, continuity, and holes.

 When topologists talk about shapes, they don't get hung up on the little details. Instead, they like to think of shapes as if they were made of rubber. You can stretch, bend, and twist a rubber shape any way you like, but you can't tear or glue it. So, in the world of topology, a rubber ball and a rubber cube are considered the same, because you can morph one into the other without tearing or gluing.

 However, within this realm of seemingly fluid forms, certain characteristics remain constant, and it's these properties that mathematicians use to classify and understand shapes. For example, loops, holes and handles are some of the fundamental characteristics that define shapes in topology.

 In topology, a loop is basically a closed curve, like a circle or an ellipse, within a shape. Holes, on the other hand, are the empty spaces within a shape that are completely surrounded by the shape itself. Think of the hole in a doughnut or the hollow center of a tube. Handles are a bit trickier to visualize, but imagine taking a coffee mug and focusing on the handle – that's what topologists call a handle.

 Next, we have to understand what Euclidean space represents in mathematics. In general, Euclidean space can be thought of as a flat space that has straight lines, angles, and distances that behave the way we learn in school. It's like the space we see around us, where objects move in straight lines and parallel lines stay the same distance apart.

 The Poincare Conjecture is a question about a specific type of shape known as a simply connected, closed 3-dimensional manifold. To break down this term, let's first discuss what a manifold is. In general, a manifold is a topological space that resembles Euclidean space on a small scale (locally). Imagine the surface of a sphere or the surface of a doughnut; both are examples of manifolds. A 3-dimensional manifold, as you might guess, is a 3-dimensional version of this concept, such as the space we inhabit. In more specific mathematical terms,

a manifold is defined as a special set that is locally homeomorphic to Euclidean space.

 This might sound complex and abstruse, but bear with me as I try to simplify the above statement in a colloquial manner.  

 To better explain manifolds, let me illustrate an analogy. Imagine you have a large piece of paper that you have folded into the shape of a sphere. Now, presume that you're a little ant walking on the surface of the sphere. From your perspective (the ant), the paper looks totally flat, but if you zoom out, you'll see that the paper is actually curved and shaped like a sphere. This is kind of like what a manifold is in mathematics. So, essentially, a manifold is a space that looks flat and simple when you zoom in really close, but when you zoom out, you'll see that it's actually curved or twisted in some way. Just like how the ant on the paper doesn't realize that it's walking on a sphere because it's too small to see the bigger picture, we often can't tell that a space is a manifold because we're too zoomed in.

Figure 4: A simple front-view diagram illustrating the analogy I presented. (Not to scale)

 Now, let's suppose that Figure 4 occurs in 2-dimensional space. Then, in the analogy above, the ant is perceiving the surface as a 1-dimensional line segment. But in reality, the zoomed out perspective suggest that the ant is walking on an infinitely large amount of infinitesimal 1-dimensional line segments that wrap around to form a curved shape that represents a 2-dimensional circle. So, when you zoom in infinitely into a manifold, the local space observed usually represents a $n-1$ dimension of the shape perceived, where $n$ is the dimensions of the actual shape. Thus, in the mathematics of topology, an object in $n$-dimensions will be the surface of an object with $n+1$ dimensions.

 Although these analogies are my best attempts to simplify manifolds, they are not entirely precise to the definition aforementioned. This is because the definition presented represents an abstract idea because it does not lay out explicit references to how a manifold can be viewed geometrically, rather only displays the specific sets or conditions that form a manifold. However, for this article, a cursory understanding of manifolds should be sufficient to understand the subsequent sections.

 Now that we have a general idea of what Euclidean space and manifolds are, we can further explore the concepts related to the Poincaré Conjecture. To do this, I'll begin with delving into more specific definitions and mathematical concepts that I believe are essential for a sufficient understanding of this conjecture:

  1. Simply Connected

In topology, the term "simply connected" refers to a shape that has no holes or handles. To be more precise, a shape is simply connected if every loop inside the shape can be continuously shrunk to a single point without leaving the shape. For example, a sphere is simply connected because you can draw a loop on its surface and continuously shrink it to a point, all the while keeping the loop on the surface.

Figure 5: A sphere is simply connected because any loop drawn on the surface can be shrunk to a single point without leaving the shape.

 To better understand this, try to imagine drawing a circle on a blank piece of paper. Then, visualize drawing another circle with smaller radius inside the circle you drew. Repeat this process and you'll observe that as the radius infinitely decreases, it will converge at one singular point. Note that it does not have to be a circle, it can be any shape as long as it does not have a hole. For example, a square, rectangle and even a triangle will showcase the same property. I used a circle as an example only because of its simplicity.

Figure 6: A simple diagram showing how a circle is a simply connected shape through the analogy I presented. The red line represents the actual shape of the circle (loop) drawn, while the green animation shows that any loop within the circle can eventually shrink and converge at one point (in this case, the center).

  Now, imagine doing the same thing, but this time replace the circle with the shape of a 2-dimensional doughnut. In mathematics, this shape is known as an annulus, which is a 2-dimensional manifold.

Figure 7: The shape on the left is an annulus. The shape on the right is the same annulus, but with animations that shows a loop continuously shrinking down to a single point. The green animation displays the section of the shrinking that remains within the boundaries of the annulus, whereas the orange animation shows that the loop has to leave the boundaries of the annulus in order to shrink to become a single point. 

 As seen in Figure 7, an annulus is therefore not simply connected as any loop inside the shape can only be shrunk to a single point by leaving the boundaries of the shape. However, this is far from a well-defined proof as it does not take into account the mathematical concepts surrounding this common topological property. You can click on the button below to view the article I wrote that covers the mathematical proof behind why an annulus is not simply connected.

2.  Closed

In the context of topology and manifolds, "closed" means that the manifold is compact and has no boundary. Compactness means that the shape is finite in size and can be contained within a finite region of space. No boundary implies that there are no edges or abrupt endings to the shape. For example, the surface of a sphere is a closed 2-dimensional manifold because it is finite in size and has no boundaries or edges.

Figure 8: A simple diagram showing a closed 2-dimensional manifold on a 3-dimensional sphere.

 With these definitions in mind, I can now state the Poincaré Conjecture:

The Poincaré Conjecture posits that every simply connected, closed 3-dimensional manifold is topologically equivalent to a 3-dimensional sphere (3-sphere).

 In other words, if you have a 3-dimensional shape without holes or handles that is finite and has no boundaries, can you continuously deform it into a 3-dimensional sphere without tearing or gluing? The conjecture, proposed by Henri Poincaré in 1904, suggests that the answer is yes.

 As an extra note, it is very important to understand that a 3-sphere in topology forms the boundary of a 4-dimensional ball. This means that the 3-sphere actually exists within the realm of 4-dimensional space. A 3-sphere is very difficult to visualize directly due to our 3D spatial intuition. But mathematically, it has a well-defined construction as a 4D surface. To further clear things up, a 3-sphere is nothing like a regular sphere ball you would see in real life.

Figure 9: The shape on the left is a 2-sphere and the shape on the right is a 3-sphere.

 Another essential concept to understand is that the entire observable universe that we live in exists within the 2-sphere. In other words, the 2-sphere is the shape of the universe.

Discover Grigori Perelman: The Enigmatic Mastermind Who Solved the Century-Old Poincaré Conjecture

Grigori Yakovlevich Perelman is a renowned Russian mathematician who gained international recognition for his groundbreaking work in the field of topology. Born in 1966, Perelman's exceptional mathematical talent became evident early in his life. He is best known for his monumental proof of the Poincaré Conjecture, a century-old problem in topology. His consummate affinity for mathematics allowed him to creatively merge different techniques invented in the past with his own innovative ideas that eventually lead to the proof of the Poincaré Conjecture.

Figure 10: Perelman in 1993.

 Recognized for his mathematical talent at a young age, he was enrolled in an after-school program led by Sergei Rukshin and later attended the prestigious Leningrad Secondary School #239, focusing on advanced mathematics and physics. In 1982, Perelman won a gold medal with a perfect score at the International Mathematical Olympiad as a member of the Soviet Union team. He then studied at the School of Mathematics and Mechanics at Leningrad State University without undergoing admission examinations.

Figure 11: 1982 International Mathematics Olympiad results displaying the three individuals who achieved perfect scores including Grigori Perelman.

 Perelman completed his PhD in 1990 and began working at the Leningrad Department of Steklov Institute of Mathematics, under the guidance of esteemed mathematicians Aleksandr Aleksandrov and Yuri Burago. In the early 1990s, he held research positions at various American universities, including New York University, Stony Brook University, and the University of California, Berkeley.

 Among his many accomplishments during this period, Perelman's work on Aleksandrov's spaces of curvature and manifolds with lower bounds on Ricci curvature stand out as truly groundbreaking. His profound understanding of the underlying structures of the universe culminated in the proof of the soul conjecture, a feat that earned him the esteemed Young Mathematician Prize of the St. Petersburg Mathematical Society in 1991.

 Perelman's magnum opus, however, came in 2003 when he unveiled his proof of the Poincaré Conjecture, a problem that had stumped mathematicians for over a century. By deciphering the enigma of three-dimensional spaces, Perelman etched his name into the annals of mathematical history. Surprisingly, in a display of humility and decorous principle, he declined both the Fields Medal and the Clay Millennium Prize in which he was awarded for this proof. I am unable to provide the exact context for why he chose to reject the awards, but Perelman himself stated that his contribution was no greater than that of Richard Hamilton, a fellow mathematician that greatly contributed to the field of topology and whose invention, the Ricci flow, laid the foundations for Perelman's proof of the Poincaré conjecture.

Figure 12: Richard Streit Hamilton pictured in 1982

 It is also believed that Perelman declined the awards due to an ethical breach by fellow field medalist winner, Shing-Tung Yau who, according to Perelman, allegedly attempted to downplay Perelman's role in the proof and play up the work of Cao and Zhu. Perelman was quoted in a 2006 article by The New Yorker,

"I can't say I'm outraged. Other people do worse. Of course, there are many mathematicians who are more or less honest. But almost all of them are conformists. They are more or less honest, but they tolerate those who are not honest. It is not people who break ethical standards who are regarded as aliens. It is people like me who are isolated."

 Perelman quit his prestigious job at Steklov Institute in 2005. In fact, according to several sources close to Perelman, he currently finds mathematics a painful topic to discuss; by 2010, some even said that he had entirely abandoned mathematics. When asked about this in an open letter by german artist Ursula Neugebauer in 2010, he was quoted as writing,

"I think that in Mathematics beauty is the essential criteria, as in music and art. But the understanding of beauty changes itself. According to the Platonist philosophy, the world of mathematics exists, and we discover it. However, I would prefer to say that mathematics doesn't really exist. It is not so that we open the door to a room, and you look into another room and so on. There is no door and there is no room. We create the door and we open it. We create the room and we look inside."

Perelman's Proof of the Poincaré Conjecture


It is practically impossible for me to explain Perelman's proof of the Poincaré Conjecture as it involves prodigiously complex algebraic topology that is far too above my marginal understanding of mathematics at this point of time. However, I can attempt to provide a cursory look-through at his proof and the steps he took to solve this unimaginably tough problem.

Ricci Flow: The Foundation of Perelman's Proof


Ricci flow, introduced by Richard S. Hamilton in 1982, is a process that smooths out the geometry of a manifold by evolving its metric tensors (a mathematical function that describes distances on the manifold) according to a specific equation. The idea behind Ricci flow is analogous to the process of heat diffusion, where heat spreads evenly across a surface over time. Similarly, Ricci flow redistributes the curvature of a manifold, ultimately smoothing out its geometry. The formula is written as $\frac{d}{{dt}}g_{ij} (t) =  - 2R_{ij}$, where $g_{ij} (t)$ is the Riemannian metric tensor function. The reason the equation does not use regular metric tensors is because it deals with smooth metrics on a closed manifold.

Figure 13: Intermediate stages of Ricci flow on a 2-manifold

 In other words, a regular metric tensor is used in flat, Euclidean spaces, like the 2D plane or 3D space we're familiar with in everyday life. It measures distances and angles using the standard Euclidean geometry, where straight lines remain straight, and angles are preserved. As an analogy, imagine a flat map of a city. A regular metric tensor would help you measure the distances between different locations and the angles between streets, just as you would expect in a flat, Euclidean space.

 Conversely, a Riemannian metric tensor on the other hand is used in curved spaces, called manifolds, which can be thought of as smooth, curved generalizations of Euclidean spaces. It measures distances and angles while taking into account the curvature of the manifold. As an analogy,  imagine a map of the Earth's surface. Since Earth is a sphere (a curved space), using a flat map to measure distances and angles would lead to inaccuracies. A Riemannian metric tensor would help you measure distances and angles on the Earth's surface more accurately by taking its curvature into account.

 Before continuing, it's essential that we understand the mathematical concepts surrouding curvatures in topology.

 Curvature is a measure of how much a shape bends or deviates from being flat or straight. For example, a flat piece of paper has zero curvature, while a sphere has positive curvature.

 Now, when we deal with curvature in higher-dimensional spaces or more complex shapes, we need tools to describe and analyze this curvature. In differential geometry, there are three important measures of curvature: the Riemann curvature tensor, the Ricci curvature tensor, and the scalar curvature. Each of these measures captures different aspects of curvature, and they're all related to each other.

1. Riemann curvature tensor:

Think of the Riemann curvature tensor as the most detailed, comprehensive measure of curvature. It captures the full range of information about how a shape bends in every possible direction. In real-life terms, imagine you're standing on a hilly landscape. The Riemann curvature tensor tells you not only how steep the hills and valleys are in every direction, but also how those slopes change as you move around.

 The Riemann curvature tensor is a multi-dimensional array of numbers that represents all of this information. It's quite complex, but it gives you the most complete picture of the curvature of the space you're in.

2. Ricci curvature tensor:

The Ricci curvature tensor is a simplified version of the Riemann curvature tensor. Instead of capturing all the detailed information about curvature in every direction, the Ricci curvature tensor focuses on the average curvature.

 Going back to the hilly landscape analogy, the Ricci curvature tensor tells you the average steepness of the hills and valleys. It doesn't give you all the details about how the slopes change in every direction, but it gives you a good sense of the overall "bumpiness" of the landscape.

 The Ricci curvature tensor is a multi-dimensional array of numbers, like the Riemann curvature tensor, but it's simpler and easier to work with. It's often used in applications where the full complexity of the Riemann curvature tensor isn't needed.

3. Scalar curvature:

Scalar curvature is the simplest measure of curvature, boiling everything down to a single number. It represents the overall average curvature of a shape or space. This means that a lot of the detailed information about how the shape bends in various directions is lost, but in return, you get a single, easy-to-understand measure of curvature.

 In the hilly landscape analogy, the scalar curvature gives you a single number that represents the overall "bumpiness" of the landscape. It doesn't tell you anything about the slopes in specific directions or how they change, but it gives you a general sense of how curved the space is.

 Moving ahead, the Ricci flow equation also implies that the curvature of the manifold changes according to another equation,  $R_{m_t }  =  \vartriangle R_m  + Q$, where $\vartriangle R_m$ is Laplacian (a measure of how much a function deviates from a constant), and $Q$ represents a quadratic expression of the curvatures.

 Hamilton showed that the scalar curvature ($R$) can be described as a single value that illustrates the overall curvature of the manifold and satisfies the equation $R_t  = \vartriangle R + 2|Ric|^2$, where $|Ric|$ represents the magnitude of the Ricci curvature tensor. Thus, Hamilton concluded that by the maximum principle (a mathematical tool used to study the behavior of certain functions), the minimum value of the scalar curvature does not decrease over time.

 Let me present an analogy so that you can understand this abstruse concept a little better. Imagine the manifold as a wrinkled bedsheet. The Ricci flow equation describes how the wrinkles (curvature) smooth out over time, $t$. The scalar curvature ($R$) gives an overall measure of how wrinkled the bedsheet is.

 Hamilton also showed that the Ricci flow preserves the positivity of the Ricci tensor in 3D and the curvature operator in all dimensions. This means that as the manifold evolves, its curvature remains positive or zero. Additionally, the eigenvalues (a measure of how much the curvature stretches or compresses the manifold) becomes more similar (or pinched) as the curvature increases.

 Going back to the bedsheet example, positivity means that as you smooth out the wrinkles, you don't create any new wrinkles or folds. Pinching refers to the fact that as the bedsheet becomes smoother, the differences in the amount of stretching or compression across the sheet become smaller.

Ricci Flow with Surgery on Three Manifolds


In the 1980s, Richard S. Hamilton laid the scaffoldings for the idea of evolving a closed simply-connected manifold into a 3-dimensional sphere by continuously deforming the manifold via his Ricci flow equation. However, singularities (regions where curvature becomes unbounded) can develop during the evolution. To address this issue, Hamilton proposed performing surgery on the manifold by cutting out regions around the singularities and replacing them with "standard" pieces. This process is called "Ricci flow with surgery."

 Unfortunately, one of the challenges in Hamilton's original approach was the precise control and analysis of the surgery process. In particular, there were concerns about the possibility of the equation returning an infinite number of surgeries occurring in a finite amount of time, which would make the process poorly defined. And this was the barrier that stumped Hamilton from crafting his own detailed definition for both Thurston's Geometrization Conjecture and the Poincaré Conjecture.

 This barrier in topology was finally broken by Grigori Perelman in 2002 and 2003 when he introduced a number of key innovations and techniques to better understand and control the surgery process. Some of his most important breakthroughs are as follows,

Reduced Volume and Perelman's Entropy:

Perelman introduced a new quantity called the "reduced volume" to better understand the behavior of the Ricci flow. The reduced volume is an invariant quantity that remains constant under the Ricci flow of a 3-manifold.  

 To better understand the concept of reduced volume, picture a lengthy and complex border between two countries. The total length of this border represents the "unreduced volume." By simplifying the border and straightening out its inlets and peninsulas, we create a shorter border, which we call the "reduced volume." This demonstrates that by simplifying a shape and smoothing its curves and folds, we can decrease its volume.

 In the context of a 3-manifold, the reduced volume remains constant under Ricci flow. This is because Ricci flow is a continuous process that smooths out the manifold's shape, thus reducing its volume. However, it cannot go below a minimum value (for a 3-manifold, the maximum reduced volume corresponds to a 3-sphere). With this, Perelman demonstrated that, as a result, the reduced volume must always be constant for Ricci flow in 3-manifolds.

 Perelman also defined a function called "Perelman's entropy (referred to as geometric entropy in his papers)," which helps us understand how the geometry of the manifold changes as we apply the Ricci flow. Perelman's entropy is a scalar quantity that shows the measure of how disordered or random a shape is in its configuration space. It quantifies how spread out and tangled its possible configurations are. The Perelman entropy is defined by the following function:

$$\mathcal{W}(g,f,\tau) = \int_M \left[ \tau \left( -(4\pi\tau)^{-\frac{n}{2}}e^{-f} \right)^{\frac{n}{2}} R + \left| \nabla f - \frac{\nabla \tau}{2\tau} \right|^2 \left( -(4\pi\tau)^{-\frac{n}{2}}e^{-f} \right)^{\frac{n}{2}} \right] dV_g,$$

where $g$ is the metric of the manifold $M$, $f$ is a scalar function, $\tau$ is a positive parameter, $R$ is the scalar curvature, and $n$ is the dimension of the manifold.

 To illustrate how this works, think of a messy, disorganized room. The entropy in this scenario would correspond to all the possible ways that the various items in the room could have been arranged to create that same mess. A higher entropy mess has many more possible arrangements, indicating greater disorder.

 Now, you might be thinking that this means Ricci flow will always decrease the Perelman entropy as it smoothens the shape, but this is not always true. The reason is that Ricci flow can create singularities, which are points where the curvature becomes unbounded. These singularities can increase the Perelman entropy, even though the flow smoothens the geometry elsewhere. Thus, in order to overcome this, Perelman modified Hamilton's Ricci flow with surgery to deal with these singularities, ensuring that the 3-manifold will have entropy that behaves as a monotonic function that decreases during Ricci flow.

Canonical Neighborhoods:

Perelman introduced the concept of "canonical neighborhoods" to help classify the structure of the manifold near singularities (points where the Ricci flow breaks down) and guide the surgery process.

 A canonical neighborhood is a small region surrounding a singularity where Perelman was able to analyze the geometry and topology in a controlled way. Specifically, he showed that the geometry in these neighborhoods could be normalized into a standard "canonical" form. Understanding the topology of these canonical neighborhoods then guided Perelman in performing delicate surgeries to remove the singular region and continue the Ricci flow.

 To illustrate, imagine you have a large, complicated jigsaw puzzle that is partially solved, but contains some missing pieces as well as extra pieces that don't belong. Now suppose this isn't an ordinary puzzle - it has been designed so that there are extra pieces provided beyond what is necessary. Specifically, this 100-piece puzzle has been given to you with 200 pieces total. It is also worth nothing that in this puzzle,  the extra pieces can be cut and reshaped. The goal is to solve the puzzle using exactly 100 pieces, disregarding the extra pieces provided. In other words, you need to determine which pieces are essential to the solution and which can be cast aside.

 Then, when you start doing the puzzle, you realize that there are many odd gaps when trying to solve it with 100 pieces. In this scenario, a canonical neighborhood is like a versatile template that analyzes these gaps in the puzzle and defines the shape of the gaps in a detailed manner, allowing you to construct the actual pieces to fit into those gaps. But at the same time, we can't just magically create new pieces out of nothing. So, we then have to carefully construct the puzzle pieces that matches that template, by cutting it out of the extra pieces we already have available in the solved part of the puzzle nearby.

 In this context, the entire puzzle with its 200 pieces represents an arbitrary 3-manifold, and each of those 200 pieces represents specific topological structures that could be added or removed to continue simplifying its geometry.

Gap Lemma and Surgery Control:

Perelman proved a "gap lemma," which is a result that states that the geometry of the manifold before and after surgery is well-separated. This is important for controlling the surgery process and showing that surgeries can be performed in a well-defined way.

 To illustrate, imagine that you're trying to cut a cake into two pieces. You want to make sure that the two pieces are roughly equal in size and that the cut is smooth. The gap lemma is like a guarantee that you can make the cut in a way that separates the cake perfectly into two pieces without destroying the overall shape of the cake.

No local collapsing theorem:

Perelman showed that if the initial manifold satisfies a certain curvature condition, then the manifold does not collapse locally under the Ricci flow. This result was important for understanding the long-time behavior of Ricci flow with surgery.

 For instance, suppose that you're trying to build a tower out of blocks. If the blocks are too weak or they're not stacked properly in a well-defined manner, the tower might collapse. Similarly, Perelman's result ensures that the geometry of the manifold is stable under the Ricci flow, so we don't have to worry about the manifold "collapsing" or becoming unstable over time.

Finite time extinction:

As presented earlier, one of the main reasons Hamilton's Ricci flow with surgery was not well-defined was due to the possibility of an infinite number of surgeries occurring in a finite amount of time, which simply does not make any sense. Thus, in order to exterminate this events, Perelman introduced "finite-time extinction".

 Perelman showed that after a finite number of surgeries, the manifold would reach a state where no more surgeries are required, and the manifold would evolve under Ricci flow to become extinct in a finite amount of time. This extinction implies that the manifold has reached a state where it is topologically equivalent to a sphere or a collection of spheres.

 To illustrate in a more pellucid manner, suppose that you have a lump of clay that represents a 3-dimensional manifold (a smooth surface without any holes or discontinuities). You want to reshape this lump of clay to obtain a collection of spheres, which are the simplest possible 3D shapes.

 In Hamilton's Ricci flow, think of it as the natural process that smooths out the clay over time, like water flowing over a rock, eroding it into a simpler shape. Ricci flow is a mathematical process that gradually changes the geometry of the manifold, so it becomes more uniform. However, sometimes the smoothing process isn't enough to simplify the clay into spheres. You might need to perform a "surgery" to remove a bump or a dent in the clay. In mathematical terms, this involves cutting out a specific part of the manifold and sewing it back together in a simpler way. And this was where Hamilton experienced issues with his solution, as the big problem was that, in some cases, Hamilton's solution suggested that you might need to perform an infinite number of surgeries in a finite amount of time.

 Fortunately, Perelman introduced the concept of "finite-time extinction" to address this issue. Imagine that after performing a certain number of surgeries on the clay, you reach a point where the smoothing process (Ricci flow) can take over and finish the job without needing any more surgeries. The clay is now guaranteed to become a collection of spheres in a finite amount of time.

 With this, Perelman successfully addressed the problems involving Hamilton's Ricci flow with surgery and proved that for any simply connected closed 3-dimensional manifold, you can reach a state where no more surgeries are needed, and the manifold will evolve under Ricci flow to become a sphere or a collection of spheres.

Summary

To reiterate, the Poincaré conjecture posits that every simply connected, closed 3-dimensional manifold is topologically equivalent to a 3-dimensional sphere.

In summary, Perelman proofed this conjecture by:

  1. Start with any simply connected, closed 3-dimensional manifold.
  2. Apply Hamilton's Ricci flow to smooth out the manifold's geometry.
  3. During the Ricci flow process, any singularities that emerge can be dealt with through surgery, and then continue the Ricci flow process.
  4. If an event occurs where the equation requires an infinite number of surgeries in a finite time, then eliminate the event through Perelman's finite-time extinction.
  5. Continue this process, and eventually, the 3-manifold will reach a state where it is topologically equivalent to a 3-dimensional sphere.

 And just like that, an enigma that had endured for a century, confounding the most brilliant and prodigious minds in mathematics, gracefully unfurled its mystery. The Poincaré conjecture and Thurston's Geometrization conjecture, once elusive and shrouded in uncertainty, became illuminated by the triumph of human intellect.


References

  1. T., D.L.A. (1912) Jules Henri Poincaré. Toronto: s.n.
  2. Henri Poincaré (2023) Wikipedia. Wikimedia Foundation. Available at: https://en.wikipedia.org/wiki/Henri_Poincar%C3%A9.
  3. Barrow-Green, J. (1997) Poincaré and the three body problem. Providence, RI: American Mathematical Society.
  4. Hsu, J.P. and Hsu, L. (2006) “vol. 10 p. 37,” in A broader view of relativity general implications of Lorentz and Poincaré invariance. New Jersey: World Scientific Publishing.
  5. Feng, J. (2017) University of Paris - educational institutions around the world, WorldAtlas. WorldAtlas. Available at: https://www.worldatlas.com/articles/university-of-paris-educational-institutions-around-the-world.html.
  6. Poincaré Henri et al. (2022) Science and hypothesis. London: Bloomsbury Academic.
  7. Euclidean space (2023) Wikipedia. Wikimedia Foundation. Available at: https://en.wikipedia.org/wiki/Euclidean_space.
  8. Manifold (no date) from Wolfram MathWorld. Available at: https://mathworld.wolfram.com/Manifold.html.
  9. Keng, B. (2018) Manifolds: A gentle introduction, Bounded Rationality. Github. Available at: https://bjlkeng.github.io/posts/manifolds/.
  10. Thurston, W.P. and Weeks, J.R. (1984) The mathematics of three-dimensional manifolds - cornell university, Scientific American Inc. Cornell University. Available at: https://pi.math.cornell.edu/~mathclub/Media/math-of-3-dimensional-manifolds.pdf.
  11. Simply connected space - université de montréal (no date). Université de Montréal. Available at: https://www.astro.umontreal.ca/~bergeron/PHY2345/ConnexiteSimple.pdf.
  12. Faust, G.H. (2014) Prove that an annulus is not simply connected?, Mathematics Stack Exchange. StackExchange. Available at: https://math.stackexchange.com/questions/695948/prove-that-an-annulus-is-not-simply-connected.
  13. Fundamental group (2023) Wikipedia. Wikimedia Foundation. Available at: https://en.wikipedia.org/wiki/Fundamental_group.
  14. Closed set (2023) Wikipedia. Wikimedia Foundation. Available at: https://en.wikipedia.org/wiki/Closed_set.
  15. Lizvullok (2018) 2-spheres and 3-spheres, YouTube. YouTube. Available at: https://www.youtube.com/watch?v=QKTvqAjwXe0.
  16. Grigori Perelman (2023) Wikipedia. Wikimedia Foundation. Available at: https://en.wikipedia.org/wiki/Grigori_Perelman.
  17. International mathematical olympiad (1982) International Mathematical Olympiad. Available at: https://www.imo-official.org/year_individual_r.aspx?year=1982&column=total&order=desc.
  18. Moseman, A. (2019) Brilliant & reclusive Russian mathematician doesn't need your prize money, Discover Magazine. Discover Magazine. Available at: https://www.discovermagazine.com/the-sciences/brilliant-and-reclusive-russian-mathematician-doesnt-need-your-prize-money.
  19. Sylvia Nasar, D.G. (2006) The Poincaré clash, The New Yorker. Available at: https://www.newyorker.com/magazine/2006/08/28/manifold-destiny.
  20. Richard S. Hamilton (2023) Wikipedia. Wikimedia Foundation. Available at: https://en.wikipedia.org/wiki/Richard_S._Hamilton.
  21. Neugebauer, U. (2010) Das größte Glück, welches vorstellbar ist, Ursula Neugebauer. Available at: http://www.ursula-neugebauer.de/?p=184.
  22. Ricci flow (2023) Wikipedia. Wikimedia Foundation. Available at: https://en.wikipedia.org/wiki/Ricci_flow.
  23. Anderson, M.T. (2004) Geometrization of 3-manifolds via the Ricci flow. State University of New York. Available at: https://www.ams.org/journals/notices/200402/fea-anderson.pdf.
  24. Riemann curvature tensor (2023) Wikipedia. Wikimedia Foundation. Available at: https://en.wikipedia.org/wiki/Riemann_curvature_tensor.
  25. Ricci curvature (2023) Wikipedia. Wikimedia Foundation. Available at: https://en.wikipedia.org/wiki/Ricci_curvature.
  26. Perelman, G. (2002) The entropy formula for the Ricci flow and its geometric applications, arXiv.org. arXiv. Available at: https://arxiv.org/abs/math/0211159.
  27. Perelman, G. (2003) Ricci flow with surgery on three-manifolds, arXiv.org. Available at: https://arxiv.org/abs/math/0303109.
  28. Perelman, G. (2003) Finite extinction time for the solutions to the Ricci flow on certain three-manifolds, arXiv.org. Available at: https://arxiv.org/abs/math/0307245.