Search This Blog

Friday, July 27, 2018

Is Fermat's Last Theorem Stated Too Narrowly?

I am not a mathematician, but I do enjoy math and numbers, and can't resist playing around a little.  One of my "toys" has been Fermat's Last Theorem.  For a broader explanation of this theorem and the proof of it, see https://en.wikipedia.org/wiki/Fermat's_Last_Theorem, which I will summarize here with the first paragraph:

"In number theory, Fermat's Last Theorem (sometimes called Fermat's conjecture, especially in older texts) states that no three positive integers a, b, and c satisfy the equation an + bn = cn for any integer value of n greater than 2. The cases n = 1 and n = 2 have been known to have infinitely many solutions since antiquity."

As you might know, the proof of this theorem was, and still is, an historic mystery. Fermat claimed to have proved it himself around 1637, but never provided the proof, only a claim "in the margin of a copy of Arithmetica where he claimed he had a proof that was too large to fit in the margin." Yet if Fermat truly did have a valid proof, he never wrote it out, even until his death in 1665.


It was only in 1994. 357 years later after Fermat's cryptic and unproved claim that the mathematician Andrew Wiles finally presented the first proof of the theorem which was accepted by the mathematical community as correct and flawless (please don't ask me to explain it, however!).  This in itself is astonishing enough, but what is more astonishing is the fact that Wiles used mathematical techniques which did not exist in Fermat's time.  Thus, if Fermat really did supply a proof, he did so using techniques much simpler -- techniques which have evaded the most brilliant math minds for over three hundred years!  One can only conclude that either Fermat erred in his claim, or a straightforward proof has been eluding such minds to this day.

Again, I warn I am no mathematician, but since Andrew Wiles' proof I have wondered if Fermat's theorem can be expanded upon.  The theorem may follow a pattern, in which the number of integers being added must equal the exponent itself.  For example, for 32 + 42 = 52, the simplest example of the theorem, that number is two.

What about three?  It turns out there is an example here, one that I stumbled upon quite accidentally:  33 + 43 + 53 = 63. Now I suspect there are more examples for three, just as there are for two, although I have not tried to find any of them.  Nor have I attempted to find examples for integers greater than three, though I suspect they exist.

You probably see where I'm going with this.  Is it a reasonable conjecture that, for all (positive?) integers n, there are sums of n integers raised to the n'th power which will yield another integer raised to n?  Furthermore, is it an iff (if and only if) condition for all integers greater than one?

Of course, I've no idea.  I certainly unable to prove or disprove such a conjecture.  But it does seem worthy to make the attempt (if no one has done so yet).  Perhaps one of you can do so, or at least has some logical input on the question.  I'll leave it at that.




Measurement problem

From Wikipedia, the free encyclopedia
 
The measurement problem in quantum mechanics is the problem of how (or whether) wave function collapse occurs. The inability to observe this process directly has given rise to different interpretations of quantum mechanics, and poses a key set of questions that each interpretation must answer. The wave function in quantum mechanics evolves deterministically according to the Schrödinger equation as a linear superposition of different states, but actual measurements always find the physical system in a definite state. Any future evolution is based on the state the system was discovered to be in when the measurement was made, meaning that the measurement "did something" to the system that is not obviously a consequence of Schrödinger evolution.

To express matters differently (to paraphrase Steven Weinberg[1][2]), the Schrödinger wave equation determines the wave function at any later time. If observers and their measuring apparatus are themselves described by a deterministic wave function, why can we not predict precise results for measurements, but only probabilities? As a general question: How can one establish a correspondence between quantum and classical reality?[3]

Schrödinger's cat

The best known example is the "paradox" of the Schrödinger's cat. A mechanism is arranged to kill a cat if a quantum event, such as the decay of a radioactive atom, occurs. Thus the fate of a large scale object, the cat, is entangled with the fate of a quantum object, the atom. Prior to observation, according to the Schrödinger equation, the cat is apparently evolving into a linear combination of states that can be characterized as an "alive cat" and states that can be characterized as a "dead cat". Each of these possibilities is associated with a specific nonzero probability amplitude; the cat seems to be in some kind of "combination" state called a "quantum superposition". However, a single, particular observation of the cat does not measure the probabilities: it always finds either a living cat, or a dead cat. After the measurement the cat is definitively alive or dead. The question is: How are the probabilities converted into an actual, sharply well-defined outcome?

Interpretations

The Copenhagen interpretation is the oldest and probably still the most widely held interpretation of quantum mechanics.[4][5][6] Most generally it posits something in the act of observation which results in the collapse of the wave function. According to the von Neumann–Wigner interpretation the causative agent in this collapse is consciousness.[7] How this could happen is widely disputed. Hugh Everett's many-worlds interpretation attempts to solve the problem by suggesting there is only one wave function, the superposition of the entire universe, and it never collapses—so there is no measurement problem. Instead, the act of measurement is simply an interaction between quantum entities, e.g. observer, measuring instrument, electron/positron etc., which entangle to form a single larger entity, for instance living cat/happy scientist. Everett also attempted to demonstrate the way that in measurements the probabilistic nature of quantum mechanics would appear; work later extended by Bryce DeWitt.

De Broglie–Bohm theory tries to solve the measurement problem very differently: the information describing the system contains not only the wave function, but also supplementary data (a trajectory) giving the position of the particle(s). The role of the wave function is to generate the velocity field for the particles. These velocities are such that the probability distribution for the particle remains consistent with the predictions of the orthodox quantum mechanics. According to de Broglie–Bohm theory, interaction with the environment during a measurement procedure separates the wave packets in configuration space which is where apparent wave function collapse comes from even though there is no actual collapse.

The Ghirardi–Rimini–Weber (GRW) theory differs from other collapse theories by proposing that wave function collapse happens spontaneously. Particles have a non-zero probability of undergoing a "hit", or spontaneous collapse of the wave function, on the order of once every hundred million years.[8] Though collapse is extremely rare, the sheer number of particles in a measurement system means that the probability of a collapse occurring somewhere in the system is high. Since the entire measurement system is entangled (via quantum entanglement), the collapse of a single particle initiates the collapse of the entire measurement apparatus.

Erich Joos and Heinz-Dieter Zeh claim that the phenomenon of quantum decoherence, which was put on firm ground in the 1980s, resolves the problem.[9] The idea is that the environment causes the classical appearance of macroscopic objects. Zeh further claims that decoherence makes it possible to identify the fuzzy boundary between the quantum microworld and the world where the classical intuition is applicable.[10][11] Quantum decoherence was proposed in the context of the many-worlds interpretation[citation needed], but it has also become an important part of some modern updates of the Copenhagen interpretation based on consistent histories.[12][13] Quantum decoherence does not describe the actual process of the wave function collapse, but it explains the conversion of the quantum probabilities (that exhibit interference effects) to the ordinary classical probabilities. See, for example, Zurek,[3] Zeh[10] and Schlosshauer.[14]

The present situation is slowly clarifying, as described in a recent paper by Schlosshauer as follows:[15]
Several decoherence-unrelated proposals have been put forward in the past to elucidate the meaning of probabilities and arrive at the Born rule ... It is fair to say that no decisive conclusion appears to have been reached as to the success of these derivations. ...
As it is well known, [many papers by Bohr insist upon] the fundamental role of classical concepts. The experimental evidence for superpositions of macroscopically distinct states on increasingly large length scales counters such a dictum. Superpositions appear to be novel and individually existing states, often without any classical counterparts. Only the physical interactions between systems then determine a particular decomposition into classical states from the view of each particular system. Thus classical concepts are to be understood as locally emergent in a relative-state sense and should no longer claim a fundamental role in the physical theory.
A fourth approach is given by objective collapse models. In such models, the Schrödinger equation is modified and obtains nonlinear terms. These nonlinear modifications are of stochastic nature and lead to a behaviour which for microscopic quantum objects, e.g. electrons or atoms, is unmeasurably close to that given by the usual Schrödinger equation. For macroscopic objects, however, the nonlinear modification becomes important and induces the collapse of the wave function. Objective collapse models are effective theories. The stochastic modification is thought of to stem from some external non-quantum field, but the nature of this field is unknown. One possible candidate is the gravitational interaction as in the models of Diósi and Penrose. The main difference of objective collapse models compared to the other approaches is that they make falsifiable predictions that differ from standard quantum mechanics. Experiments are already getting close to the parameter regime where these predictions can be tested.[16]

An interesting solution to the measurement problem is also provided by the hidden-measurements interpretation of quantum mechanics. The hypothesis at the basis of this approach is that in a typical quantum measurement there is a condition of lack of knowledge about which interaction between the measured entity and the measuring apparatus is actualized at each run of the experiment. One can then show that the Born rule can be derived by considering a uniform average over all these possible measurement-interactions.

The $2.5 trillion reason we can’t rely on batteries to clean up the grid

Fluctuating solar and wind power require lots of energy storage, and lithium-ion batteries seem like the obvious choice—but they are far too expensive to play a major role.

A pair of 500-foot smokestacks rise from a natural-gas power plant on the harbor of Moss Landing, California, casting an industrial pall over the pretty seaside town.
If state regulators sign off, however, it could be the site of the world’s largest lithium-ion battery project by late 2020, helping to balance fluctuating wind and solar energy on the California grid.

The 300-megawatt facility is one of four giant lithium-ion storage projects that Pacific Gas and Electric, California’s largest utility, asked the California Public Utilities Commission to approve in late June. Collectively, they would add enough storage capacity to the grid to supply about 2,700 homes for a month (or to store about .0009 percent of the electricity the state uses each year).

The California projects are among a growing number of efforts around the world, including Tesla’s 100-megawatt battery array in South Australia, to build ever larger lithium-ion storage systems as prices decline and renewable generation increases. They’re fueling growing optimism that these giant batteries will allow wind and solar power to displace a growing share of fossil-fuel plants.

But there’s a problem with this rosy scenario. These batteries are far too expensive and don’t last nearly long enough, limiting the role they can play on the grid, experts say. If we plan to rely on them for massive amounts of storage as more renewables come online—rather than turning to a broader mix of low-carbon sources like nuclear and natural gas with carbon capture technology—we could be headed down a dangerously unaffordable path.

Small doses

Today’s battery storage technology works best in a limited role, as a substitute for “peaking” power plants, according to a 2016 analysis by researchers at MIT and Argonne National Lab. These are smaller facilities, frequently fueled by natural gas today, that can afford to operate infrequently, firing up quickly when prices and demand are high.

Lithium-ion batteries could compete economically with these natural-gas peakers within the next five years, says Marco Ferrara, a cofounder of Form Energy, an MIT spinout developing grid storage batteries.

“The gas peaker business is pretty close to ending, and lithium-ion is a great replacement,” he says.

This peaker role is precisely the one that most of the new and forthcoming lithium-ion battery projects are designed to fill. Indeed, the California storage projects could eventually replace three natural-gas facilities in the region, two of which are peaker plants.

But much beyond this role, batteries run into real problems. The authors of the 2016 study found steeply diminishing returns when a lot of battery storage is added to the grid. They concluded that coupling battery storage with renewable plants is a “weak substitute” for large, flexible coal or natural-gas combined-cycle plants, the type that can be tapped at any time, run continuously, and vary output levels to meet shifting demand throughout the day.

Not only is lithium-ion technology too expensive for this role, but limited battery life means it’s not well suited to filling gaps during the days, weeks, and even months when wind and solar generation flags.

This problem is particularly acute in California, where both wind and solar fall off precipitously during the fall and winter months. Here’s what the seasonal pattern looks like:
If renewables provided 80 percent of California electricity – half wind, half solar – generation would fall precipitously beginning in the late summer.
Clean Air Task Force analysis of CAISO data
This leads to a critical problem: when renewables reach high levels on the grid, you need far, far more wind and solar plants to crank out enough excess power during peak times to keep the grid operating through those long seasonal dips, says Jesse Jenkins, a coauthor of the study and an energy systems researcher. That, in turn, requires banks upon banks of batteries that can store it all away until it’s needed.

And that ends up being astronomically expensive.

California dreaming

There are issues California can’t afford to ignore for long. The state is already on track to get 50 percent of its electricity from clean sources by 2020, and the legislature is once again considering a bill that would require it to reach 100 percent by 2045. To complicate things, regulators voted in January to close the state’s last nuclear plant, a carbon-free source that provides 24 percent of PG&E’s energy. That will leave California heavily reliant on renewable sources to meet its goals.

The Clean Air Task Force, a Boston-based energy policy think tank, recently found that reaching the 80 percent mark for renewables in California would mean massive amounts of surplus generation during the summer months, requiring 9.6 million megawatt-hours of energy storage. Achieving 100 percent would require 36.3 million.

The state currently has 150,000 megawatt-hours of energy storage in total. (That’s mainly pumped hydroelectric storage, with a small share of batteries.)
If renewables supplied 80 percent of California electricity, more than eight million megawatt-hours of surplus energy would be generated during summer peaks.
Clean Air Task Force analysis of CAISO data.
Building the level of renewable generation and storage necessary to reach the state’s goals would drive up costs exponentially, from $49 per megawatt-hour of generation at 50 percent to $1,612 at 100 percent.

And that's assuming lithium-ion batteries will cost roughly a third what they do now.
California’s power system costs rise exponentially if renewables generate the bulk of electricity.
Clean Air Task Force analysis of CAISO data.
“The system becomes completely dominated by the cost of storage,” says Steve Brick, a senior advisor for the Clean Air Task Force. “You build this enormous storage machine that you fill up by midyear and then just dissipate it. It’s a massive capital investment that gets utilized very little.”

These forces would dramatically increase electricity costs for consumers.

“You have to pause and ask yourself: ‘Is there any way the public would stand for that?’” Brick says.

Similarly, a study earlier this year in Energy & Environmental Science found that meeting 80 percent of US electricity demand with wind and solar would require either a nationwide high-speed transmission system, which can balance renewable generation over hundreds of miles, or 12 hours of electricity storage for the whole system (see “Relying on renewables alone significantly inflates the cost of overhauling energy”).

At current prices, a battery storage system of that size would cost more than $2.5 trillion.

A scary price tag

Of course, cheaper and better grid storage is possible, and researchers and startups are exploring various possibilities. Form Energy, which recently secured funding from Bill Gates’s Breakthrough Energy Ventures, is trying to develop aqueous sulfur flow batteries with far longer duration, at a fifth the cost where lithium-ion batteries are likely to land.

Ferrara’s modeling has found that such a battery could make it possible for renewables to provide 90 percent of electricity needs for most grids, for just marginally higher costs than today’s.

But it’s dangerous to bank on those kinds of battery breakthroughs—and even if Form Energy or some other company does pull it off, costs would still rise exponentially beyond the 90 percent threshold, Ferrara says.

“The risk,” Jenkins says, “is we drive up the cost of deep decarbonization in the power sector to the point where the public decides it’s simply unaffordable to continue toward zero carbon.”

Hard problem of consciousness

From Wikipedia, the free encyclopedia

The hard problem of consciousness is the problem of explaining how and why we have qualia or phenomenal experiences—how sensations acquire characteristics, such as colors and tastes. The philosopher David Chalmers, who introduced the term "hard problem" of consciousness, contrasts this with the "easy problems" of explaining the ability to discriminate, integrate information, report mental states, focus attention, etc. Easy problems are easy because all that is required for their solution is to specify a mechanism that can perform the function. That is, their proposed solutions, regardless of how complex or poorly understood they may be, can be entirely consistent with the modern materialistic conception of natural phenomena. Chalmers claims that the problem of experience is distinct from this set, and he argues that the problem of experience will "persist even when the performance of all the relevant functions is explained".

The existence of a "hard problem" is controversial and has been disputed by philosophers such as Daniel Dennett[4] and cognitive neuroscientists such as Stanislas Dehaene.[5] Clinical neurologist and skeptic Steven Novella refers to it as "the hard non-problem".[6]

Formulation of the problem

Chalmers' formulation

In Facing Up to the Problem of Consciousness (1995), Chalmers wrote:[3]
It is undeniable that some organisms are subjects of experience. But the question of how it is that these systems are subjects of experience is perplexing. Why is it that when our cognitive systems engage in visual and auditory information-processing, we have visual or auditory experience: the quality of deep blue, the sensation of middle C? How can we explain why there is something it is like to entertain a mental image, or to experience an emotion? It is widely agreed that experience arises from a physical basis, but we have no good explanation of why and how it so arises. Why should physical processing give rise to a rich inner life at all? It seems objectively unreasonable that it should, and yet it does.
In the same paper, he also wrote:
The really hard problem of consciousness is the problem of experience. When we think and perceive there is a whir of information processing, but there is also a subjective aspect.
The philosopher Raamy Majeed noted in 2016 that the hard problem is, in fact, associated with two "explanatory targets":[7]
  1. [PQ] Physical processing gives rise to experiences with a phenomenal character.
  2. [Q] Our phenomenal qualities are thus-and-so.
The first fact concerns the relationship between the physical and the phenomenal, whereas the second concerns the very nature of the phenomenal itself. Most responses to the hard problem are aimed at explaining either one of these facts or both.

Easy problems

Chalmers contrasts the hard problem with a number of (relatively) easy problems that consciousness presents. He emphasizes that what the easy problems have in common is that they all represent some ability, or the performance of some function or behavior. Examples of easy problems include:[8]
  • the ability to discriminate, categorize, and react to environmental stimuli;
  • the integration of information by a cognitive system;
  • the reportability of mental states;
  • the ability of a system to access its own internal states;
  • the focus of attention;
  • the deliberate control of behavior;
  • the difference between wakefulness and sleep.

Other formulations

Other formulations of the "hard problem" include:
  • "How is it that some organisms are subjects of experience?"
  • "Why does awareness of sensory information exist at all?"
  • "Why do qualia exist?"
  • "Why is there a subjective component to experience?"
  • "Why aren't we philosophical zombies?"

Historical predecessors

The hard problem has scholarly antecedents considerably earlier than Chalmers, as Chalmers himself has pointed out.[9]

The physicist and mathematician Isaac Newton wrote in a 1672 letter to Henry Oldenburg:
to determine by what modes or actions light produceth in our minds the phantasm of colour is not so easie.[10]
In An Essay Concerning Human Understanding (1690), the philosopher and physician John Locke argued:
Divide matter into as minute parts as you will (which we are apt to imagine a sort of spiritualizing or making a thinking thing of it) vary the figure and motion of it as much as you please—a globe, cube, cone, prism, cylinder, etc., whose diameters are but 1,000,000th part of a gry, will operate not otherwise upon other bodies of proportionable bulk than those of an inch or foot diameter—and you may as rationally expect to produce sense, thought, and knowledge, by putting together, in a certain figure and motion, gross particles of matter, as by those that are the very minutest that do anywhere exist. They knock, impel, and resist one another, just as the greater do; and that is all they can do... [I]t is impossible to conceive that matter, either with or without motion, could have originally in and from itself sense, perception, and knowledge; as is evident from hence that then sense, perception, and knowledge must be a property eternally inseparable from matter and every particle of it.[11]
The polymath and philosopher Gottfried Leibniz wrote in 1714, as an example also known as Leibniz's gap:
Moreover, it must be confessed that perception and that which depends upon it are inexplicable on mechanical grounds, that is to say, by means of figures and motions. And supposing there were a machine, so constructed as to think, feel, and have perception, it might be conceived as increased in size, while keeping the same proportions, so that one might go into it as into a mill. That being so, we should, on examining its interior, find only parts which work one upon another, and never anything by which to explain a perception.[12]
The philosopher and political economist J.S. Mill wrote in A System of Logic (1843), Book V, Chapter V, section 3:
Now I am far from pretending that it may not be capable of proof, or that it is not an important addition to our knowledge if proved, that certain motions in the particles of bodies are the conditions of the production of heat or light; that certain assignable physical modifications of the nerves may be the conditions not only of our sensations or emotions, but even of our thoughts; that certain mechanical and chemical conditions may, in the order of nature, be sufficient to determine to action the physiological laws of life. All I insist upon, in common with every thinker who entertains any clear idea of the logic of science, is, that it shall not be supposed that by proving these things one step would be made towards a real explanation of heat, light, or sensation; or that the generic peculiarity of those phenomena can be in the least degree evaded by any such discoveries, however well established. Let it be shown, for instance, that the most complex series of physical causes and effects succeed one another in the eye and in the brain to produce a sensation of colour; rays falling on the eye, refracted, converging, crossing one another, making an inverted image on the retina, and after this a motion—let it be a vibration, or a rush of nervous fluid, or whatever else you are pleased to suppose, along the optic nerve—a propagation of this motion to the brain itself, and as many more different motions as you choose; still, at the end of these motions, there is something which is not motion, there is a feeling or sensation of colour. Whatever number of motions we may be able to interpolate, and whether they be real or imaginary, we shall still find, at the end of the series, a motion antecedent and a colour consequent. The mode in which any one of the motions produces the next, may possibly be susceptible of explanation by some general law of motion: but the mode in which the last motion produces the sensation of colour, cannot be explained by any law of motion; it is the law of colour: which is, and must always remain, a peculiar thing. Where our consciousness recognises between two phenomena an inherent distinction; where we are sensible of a difference which is not merely of degree, and feel that no adding one of the phenomena to itself would produce the other; any theory which attempts to bring either under the laws of the other must be false; though a theory which merely treats the one as a cause or condition of the other, may possibly be true.
The biologist T.H. Huxley wrote in 1868:
But what consciousness is, we know not; and how it is that anything so remarkable as a state of consciousness comes about as the result of irritating nervous tissue, is just as unaccountable as the appearance of the Djin when Aladdin rubbed his lamp in the story, or as any other ultimate fact of nature.[13]
The philosopher Thomas Nagel argued in 1974:
If physicalism is to be defended, the phenomenological features must themselves be given a physical account. But when we examine their subjective character it seems that such a result is impossible. The reason is that every subjective phenomenon is essentially connected with a single point of view, and it seems inevitable that an objective, physical theory will abandon that point of view.[14]

Relationship to scientific frameworks

Neural correlates of consciousness

Since 1990, researchers including the molecular biologist Francis Crick and the neuroscientist Christof Koch have made significant progress toward identifying which neurobiological events occur concurrently to the experience of subjective consciousness.[15] These postulated events are referred to as neural correlates of consciousness or NCCs. However, this research arguably addresses the question of which neurobiological mechanisms are linked to consciousness but not the question of why they should give rise to consciousness at all, the latter being the hard problem of consciousness as Chalmers formulated it. In "On the Search for the Neural Correlate of Consciousness", Chalmers said he is confident that, granting the principle that something such as what he terms global availability can be used as an indicator of consciousness, the neural correlates will be discovered "in a century or two".[16] Nevertheless, he stated regarding their relationship to the hard problem of consciousness:
One can always ask why these processes of availability should give rise to consciousness in the first place. As yet we cannot explain why they do so, and it may well be that full details about the processes of availability will still fail to answer this question. Certainly, nothing in the standard methodology I have outlined answers the question; that methodology assumes a relation between availability and consciousness, and therefore does nothing to explain it. [...] So the hard problem remains. But who knows: Somewhere along the line we may be led to the relevant insights that show why the link is there, and the hard problem may then be solved.[16]
The neuroscientist and Nobel laureate Eric Kandel wrote that locating the NCCs would not solve the hard problem, but rather one of the so-called easy problems to which the hard problem is contrasted.[17] Kandel went on to note Crick and Koch's suggestion that once the binding problem—understanding what accounts for the unity of experience—is solved, it will be possible to solve the hard problem empirically.[17] However, neuroscientist Anil Seth argued that emphasis on the so-called hard problem is a distraction from what he calls the "real problem": understanding the neurobiology underlying consciousness, namely the neural correlates of various conscious processes.[18] This more modest goal is the focus of most scientists working on consciousness.[17] Psychologist Susan Blackmore believes, by contrast, that the search for the neural correlates of consciousness is futile and itself predicated on an erroneous belief in the hard problem of consciousness.[19]

Integrated information theory

Integrated information theory (IIT), developed by the neuroscientist and psychiatrist Giulio Tononi in 2004 and more recently also advocated by Koch, is one of the most discussed models of consciousness in neuroscience and elsewhere.[20][21] The theory proposes an identity between consciousness and integrated information, with the latter item (denoted as Φ) defined mathematically and thus in principle measurable.[21][22] The hard problem of consciousness, write Tononi and Koch, may indeed be intractable when working from matter to consciousness.[23] However, because IIT inverts this relationship and works from phenomenological axioms to matter, they say it could be able to solve the hard problem.[23] In this vein, proponents have said the theory goes beyond identifying human neural correlates and can be extrapolated to all physical systems. Tononi wrote (along with two colleagues):
While identifying the “neural correlates of consciousness” is undoubtedly important, it is hard to see how it could ever lead to a satisfactory explanation of what consciousness is and how it comes about. As will be illustrated below, IIT offers a way to analyze systems of mechanisms to determine if they are properly structured to give rise to consciousness, how much of it, and of which kind.[24]
As part of a broader critique of IIT, Michael Cerullo suggested that the theory's proposed explanation is in fact for what he dubs (following Scott Aaronson) the "Pretty Hard Problem" of methodically inferring which physical systems are conscious—but would not solve Chalmers' hard problem.[21] "Even if IIT is correct," he argues, "it does not explain why integrated information generates (or is) consciousness."[21]

Responses

Consciousness is fundamental or elusive

Some philosophers, including David Chalmers in the late 20th century and Alfred North Whitehead earlier in the 1900s, argued that conscious experience is a fundamental constituent of the universe, a form of panpsychism sometimes referred to as panexperientialism. Chalmers argued that a "rich inner life" is not logically reducible to the functional properties of physical processes. He states that consciousness must be described using nonphysical means. This description involves a fundamental ingredient capable of clarifying phenomena that have not been explained using physical means. Use of this fundamental property, Chalmers argues, is necessary to explain certain functions of the world, much like other fundamental features, such as mass and time, and to explain significant principles in nature.

The philosopher Thomas Nagel posited in 1974 that experiences are essentially subjective (accessible only to the individual undergoing them), while physical states are essentially objective (accessible to multiple individuals). So at this stage, he argued, we have no idea what it could even mean to claim that an essentially subjective state just is an essentially non-subjective state. In other words, we have no idea of what reductivism really amounts to.[14]

New mysterianism, such as that of the philosopher Colin McGinn, proposes that the human mind, in its current form, will not be able to explain consciousness.[25]

Deflationary accounts

Some philosophers, such as Daniel Dennett[4] and Peter Hacker[26] oppose the idea that there is a hard problem. These theorists have argued that once we really come to understand what consciousness is, we will realize that the hard problem is unreal. For instance, Dennett asserts that the so-called hard problem will be solved in the process of answering the "easy" ones (which, as he has clarified, he does not consider "easy" at all).[4] In contrast with Chalmers, he argues that consciousness is not a fundamental feature of the universe and instead will eventually be fully explained by natural phenomena. Instead of involving the nonphysical, he says, consciousness merely plays tricks on people so that it appears nonphysical—in other words, it simply seems like it requires nonphysical features to account for its powers. In this way, Dennett compares consciousness to stage magic and its capability to create extraordinary illusions out of ordinary things.[27]

To show how people might be commonly fooled into overstating the powers of consciousness, Dennett describes a normal phenomenon called change blindness, a visual process that involves failure to detect scenery changes in a series of alternating images.[28] He uses this concept to argue that the overestimation of the brain's visual processing implies that the conception of our consciousness is likely not as pervasive as we make it out to be. He claims that this error of making consciousness more mysterious than it is could be a misstep in any developments toward an effective explanatory theory. Critics such as Galen Strawson reply that, in the case of consciousness, even a mistaken experience retains the essential face of experience that needs to be explained, contra Dennett.

To address the question of the hard problem, or how and why physical processes give rise to experience, Dennett states that the phenomenon of having experience is nothing more than the performance of functions or the production of behavior, which can also be referred to as the easy problems of consciousness.[4] He states that consciousness itself is driven simply by these functions, and to strip them away would wipe out any ability to identify thoughts, feelings, and consciousness altogether. So, unlike Chalmers and other dualists, Dennett says that the easy problems and the hard problem cannot be separated from each other. To him, the hard problem of experience is included among—not separate from—the easy problems, and therefore they can only be explained together as a cohesive unit.[27]

Like Dennett, Hacker argues that the hard problem is fundamentally incoherent and that "consciousness studies", as it exists today, is "literally a total waste of time":[26]
The whole endeavour of the consciousness studies community is absurd—they are in pursuit of a chimera. They misunderstand the nature of consciousness. The conception of consciousness which they have is incoherent. The questions they are asking don't make sense. They have to go back to the drawing board and start all over again.
Critics of Dennett's approach, such as Chalmers and Nagel, argue that Dennett's argument misses the point of the inquiry by merely re-defining consciousness as an external property and ignoring the subjective aspect completely. This has led detractors to refer to Dennett's book Consciousness Explained as Consciousness Ignored or Consciousness Explained Away.[4] Dennett discussed this at the end of his book with a section entitled Consciousness Explained or Explained Away?[28]

Though the most common arguments against deflationary accounts and eliminative materialism are the argument from qualia and the argument that conscious experiences are irreducible to physical states—or that current popular definitions of "physical" are incomplete—the objection follows that the one and same reality can appear in different ways, and that the numerical difference of these ways is consistent with a unitary mode of existence of the reality.[citation needed] Critics of the deflationary approach object that qualia are a case where a single reality cannot have multiple appearances. For example, the philosopher John Searle pointed out: "where consciousness is concerned, the existence of the appearance is the reality".[29]

A notable deflationary account is the higher-order theories of consciousness.[30] In 2005, the philosopher Peter Carruthers wrote about "recognitional concepts of experience", that is, "a capacity to recognize [a] type of experience when it occurs in one's own mental life", and suggested that such a capacity does not depend upon qualia.[31]

The philosophers Glenn Carruthers and Elizabeth Schier said in 2012 that the main arguments for the existence of a hard problem—philosophical zombies, Mary's room, and Nagel's bats—are only persuasive if one already assumes that "consciousness must be independent of the structure and function of mental states, i.e. that there is a hard problem". Hence, the arguments beg the question. The authors suggest that "instead of letting our conclusions on the thought experiments guide our theories of consciousness, we should let our theories of consciousness guide our conclusions from the thought experiments".[32]

In 2013, the philosopher Elizabeth Irvine pointed out that both science and folk psychology do not treat mental states as having phenomenal properties, and therefore "the hard problem of consciousness may not be a genuine problem for non-philosophers (despite its overwhelming obviousness to philosophers), and questions about consciousness may well 'shatter' into more specific questions about particular capacities".[33]

The philosopher Massimo Pigliucci distances himself from eliminativism, but he said in 2013 that the hard problem is still misguided, resulting from a "category mistake":[34]
Of course an explanation isn't the same as an experience, but that's because the two are completely independent categories, like colors and triangles. It is obvious that I cannot experience what it is like to be you, but I can potentially have a complete explanation of how and why it is possible to be you.

The source of illusion

A complete reductionistic or mechanistic theory of consciousness must include the description of a mechanism by which the subjective aspect of consciousness is perceived and reported by people. Philosophers such as Chalmers or Nagel have rejected reductionist theories of consciousness because they believe that the reports of subjective experience constitute a vast and important body of empirical evidence which is ignored by modern reductionist theories of consciousness.[9]

Dennett argued that solving the easy problem of consciousness, that is finding out how the brain works, will eventually lead to the solution of the hard problem of consciousness.[4] In particular, the solution can be achieved by identifying the stimuli and neurological pathways whose operation generates evidence of subjective experience.

Neuroscientist Michael Graziano, in his book Consciousness and the Social Brain, advocates what he calls attention schema theory, in which our perception of being conscious is merely an error in perception, held by brains which evolved to hold erroneous and incomplete models of their own internal workings, just as they hold erroneous and incomplete models of their own bodies and of the external world.[35][36]

Cognitive neuroscientist Stanislas Dehaene, in his 2014 book Consciousness and the Brain, summarized the previous decades of experimental consciousness research involving reports of subjective experience, and argued that Chalmers' "easy problems" of consciousness are actually the hard problems and the "hard problems" are based only upon ill-defined intuitions that, according to Dehaene, are continually shifting as understanding evolves:[5]
Once our intuitions are educated by cognitive neuroscience and computer simulations, Chalmers' hard problem will evaporate. The hypothetical concept of qualia, pure mental experience, detached from any information-processing role, will be viewed as a peculiar idea of the prescientific era, much like vitalism... [Just as science dispatched vitalism] the science of consciousness will keep eating away at the hard problem of consciousness until it vanishes.

Philosophy of mind

From Wikipedia, the free encyclopedia
 
A phrenological mapping[1] of the brainphrenology was among the first attempts to correlate mental functions with specific parts of the brain although it is now largely discredited.

Philosophy of mind is a branch of philosophy that studies the nature of the mind. The mind–body problem is a paradigm issue in philosophy of mind, although other issues are addressed, such as the hard problem of consciousness, and the nature of particular mental states. Aspects of the mind that are studied include mental events, mental functions, mental properties, consciousness, the ontology of the mind, the nature of thought, and the relationship of the mind to the body.

Dualism and monism are the two central schools of thought on the mind–body problem, although nuanced views have arisen that do not fit one or the other category neatly. Dualism is seen even in the Eastern tradition, in the Sankhya and Yoga schools of Hindu philosophy,[5] and Plato,[6] but its entry into Western philosophy was thanks to René Descartes in the 17th century.[7] Substance dualists like Descartes argue that the mind is an independently existing substance, whereas property dualists maintain that the mind is a group of independent properties that emerge from and cannot be reduced to the brain, but that it is not a distinct substance.[8]

Monism is the position that mind and body are not ontologically distinct entities (independent substances). This view was first advocated in Western philosophy by Parmenides in the 5th century BCE and was later espoused by the 17th century rationalist Baruch Spinoza.[9] Physicalists argue that only entities postulated by physical theory exist, and that mental processes will eventually be explained in terms of these entities as physical theory continues to evolve. Physicalists maintain various positions on the prospects of reducing mental properties to physical properties (many of whom adopt compatible forms of property dualism),[10][11][12][13][14][15] and the ontological status of such mental properties remains unclear.[14][16][17] Idealists maintain that the mind is all that exists and that the external world is either mental itself, or an illusion created by the mind. Neutral monists such as Ernst Mach and William James argue that events in the world can be thought of as either mental (psychological) or physical depending on the network of relationships into which they enter, and dual-aspect monists such as Spinoza adhere to the position that there is some other, neutral substance, and that both matter and mind are properties of this unknown substance. The most common monisms in the 20th and 21st centuries have all been variations of physicalism; these positions include behaviorism, the type identity theory, anomalous monism and functionalism.[18]

Most modern philosophers of mind adopt either a reductive or non-reductive physicalist position, maintaining in their different ways that the mind is not something separate from the body.[18] These approaches have been particularly influential in the sciences, especially in the fields of sociobiology, computer science (specifically, artificial intelligence), evolutionary psychology and the various neurosciences.[19][20][21][22] Reductive physicalists assert that all mental states and properties will eventually be explained by scientific accounts of physiological processes and states.[23][24][25] Non-reductive physicalists argue that although the mind is not a separate substance, mental properties supervene on physical properties, or that the predicates and vocabulary used in mental descriptions and explanations are indispensable, and cannot be reduced to the language and lower-level explanations of physical science.[26][27] Continued neuroscientific progress has helped to clarify some of these issues; however, they are far from being resolved. Modern philosophers of mind continue to ask how the subjective qualities and the intentionality of mental states and properties can be explained in naturalistic terms.[28][29]

Mind–body problem

The mind–body problem concerns the explanation of the relationship that exists between minds, or mental processes, and bodily states or processes.[2] The main aim of philosophers working in this area is to determine the nature of the mind and mental states/processes, and how—or even if—minds are affected by and can affect the body.
Our perceptual experiences depend on stimuli that arrive at our various sensory organs from the external world, and these stimuli cause changes in our mental states, ultimately causing us to feel a sensation, which may be pleasant or unpleasant. Someone's desire for a slice of pizza, for example, will tend to cause that person to move his or her body in a specific manner and in a specific direction to obtain what he or she wants. The question, then, is how it can be possible for conscious experiences to arise out of a lump of gray matter endowed with nothing but electrochemical properties.[18]

A related problem is how someone's propositional attitudes (e.g. beliefs and desires) cause that individual's neurons to fire and muscles to contract. These comprise some of the puzzles that have confronted epistemologists and philosophers of mind from at least the time of René Descartes.[7]

Dualist solutions to the mind–body problem

Dualism is a set of views about the relationship between mind and matter (or body). It begins with the claim that mental phenomena are, in some respects, non-physical.[8] One of the earliest known formulations of mind–body dualism was expressed in the eastern Sankhya and Yoga schools of Hindu philosophy (c. 650 BCE), which divided the world into purusha (mind/spirit) and prakriti (material substance).[5] Specifically, the Yoga Sutra of Patanjali presents an analytical approach to the nature of the mind.

In Western Philosophy, the earliest discussions of dualist ideas are in the writings of Plato who maintained that humans' "intelligence" (a faculty of the mind or soul) could not be identified with, or explained in terms of, their physical body.[6][30] However, the best-known version of dualism is due to René Descartes (1641), and holds that the mind is a non-extended, non-physical substance, a "res cogitans".[7] Descartes was the first to clearly identify the mind with consciousness and self-awareness, and to distinguish this from the brain, which was the seat of intelligence. He was therefore the first to formulate the mind–body problem in the form in which it still exists today.[7]

Arguments for dualism

The most frequently used argument in favor of dualism appeals to the common-sense intuition that conscious experience is distinct from inanimate matter. If asked what the mind is, the average person would usually respond by identifying it with their self, their personality, their soul, or some other such entity. They would almost certainly deny that the mind simply is the brain, or vice versa, finding the idea that there is just one ontological entity at play to be too mechanistic, or simply unintelligible.[8] Many modern philosophers of mind think that these intuitions are misleading and that we should use our critical faculties, along with empirical evidence from the sciences, to examine these assumptions to determine whether there is any real basis to them.[8]

Another important argument in favor of dualism is that the mental and the physical seem to have quite different, and perhaps irreconcilable, properties.[31] Mental events have a subjective quality, whereas physical events do not. So, for example, one can reasonably ask what a burnt finger feels like, or what a blue sky looks like, or what nice music sounds like to a person. But it is meaningless, or at least odd, to ask what a surge in the uptake of glutamate in the dorsolateral portion of the hippocampus feels like.

Philosophers of mind call the subjective aspects of mental events "qualia" or "raw feels".[31] There is something that it is like to feel pain, to see a familiar shade of blue, and so on. There are qualia involved in these mental events that seem particularly difficult to reduce to anything physical. David Chalmers explains this argument by stating that we could conceivably know all the objective information about something, such as the brain states and wavelengths of light involved with seeing the color red, but still not know something fundamental about the situation – what it is like to see the color red.[32]

If consciousness (the mind) can exist independently of physical reality (the brain), one must explain how physical memories are created concerning consciousness. Dualism must therefore explain how consciousness affects physical reality. One possible explanation is that of a miracle, proposed by Arnold Geulincx and Nicolas Malebranche, where all mind–body interactions require the direct intervention of God.

Another possible argument that has been proposed by C. S. Lewis[33] is the Argument from Reason: if, as monism implies, all of our thoughts are the effects of physical causes, then we have no reason for assuming that they are also the consequent of a reasonable ground. Knowledge, however, is apprehended by reasoning from ground to consequent. Therefore, if monism is correct, there would be no way of knowing this—or anything else—we could not even suppose it, except by a fluke.

The zombie argument is based on a thought experiment proposed by Todd Moody, and developed by David Chalmers in his book The Conscious Mind. The basic idea is that one can imagine one's body, and therefore conceive the existence of one's body, without any conscious states being associated with this body. Chalmers' argument is that it seems possible that such a being could exist because all that is needed is that all and only the things that the physical sciences describe about a zombie must be true of it. Since none of the concepts involved in these sciences make reference to consciousness or other mental phenomena, and any physical entity can be by definition described scientifically via physics, the move from conceivability to possibility is not such a large one.[34] Others such as Dennett have argued that the notion of a philosophical zombie is an incoherent,[35] or unlikely,[36] concept. It has been argued under physicalism that one must either believe that anyone including oneself might be a zombie, or that no one can be a zombie—following from the assertion that one's own conviction about being (or not being) a zombie is a product of the physical world and is therefore no different from anyone else's. This argument has been expressed by Dennett who argues that "Zombies think they are conscious, think they have qualia, think they suffer pains—they are just 'wrong' (according to this lamentable tradition) in ways that neither they nor we could ever discover!"[35] See also the problem of other minds.

Interactionist dualism

Portrait of René Descartes by Frans Hals (1648)

Interactionist dualism, or simply interactionism, is the particular form of dualism first espoused by Descartes in the Meditations.[7] In the 20th century, its major defenders have been Karl Popper and John Carew Eccles.[37] It is the view that mental states, such as beliefs and desires, causally interact with physical states.[8]

Descartes' famous argument for this position can be summarized as follows: Seth has a clear and distinct idea of his mind as a thinking thing that has no spatial extension (i.e., it cannot be measured in terms of length, weight, height, and so on). He also has a clear and distinct idea of his body as something that is spatially extended, subject to quantification and not able to think. It follows that mind and body are not identical because they have radically different properties.[7]

At the same time, however, it is clear that Seth's mental states (desires, beliefs, etc.) have causal effects on his body and vice versa: A child touches a hot stove (physical event) which causes pain (mental event) and makes her yell (physical event), this in turn provokes a sense of fear and protectiveness in the caregiver (mental event), and so on.

Descartes' argument crucially depends on the premise that what Seth believes to be "clear and distinct" ideas in his mind are necessarily true. Many contemporary philosophers doubt this. For example, Joseph Agassi suggests that several scientific discoveries made since the early 20th century have undermined the idea of privileged access to one's own ideas. Freud claimed that a psychologically-trained observer can understand a person's unconscious motivations better than the person himself does. Duhem has shown that a philosopher of science can know a person's methods of discovery better than that person herself does, while Malinowski has shown that an anthropologist can know a person's customs and habits better than the person whose customs and habits they are. He also asserts that modern psychological experiments that cause people to see things that are not there provide grounds for rejecting Descartes' argument, because scientists can describe a person's perceptions better than the person herself can.[41][42]

Other forms of dualism

Four varieties of dualism. The arrows indicate the direction of the causal interactions. Occasionalism is not shown.

Psychophysical parallelism

Psychophysical parallelism, or simply parallelism, is the view that mind and body, while having distinct ontological statuses, do not causally influence one another. Instead, they run along parallel paths (mind events causally interact with mind events and brain events causally interact with brain events) and only seem to influence each other.[43] This view was most prominently defended by Gottfried Leibniz. Although Leibniz was an ontological monist who believed that only one type of substance, the monad, exists in the universe, and that everything is reducible to it, he nonetheless maintained that there was an important distinction between "the mental" and "the physical" in terms of causation. He held that God had arranged things in advance so that minds and bodies would be in harmony with each other. This is known as the doctrine of pre-established harmony.[44]

Occasionalism

Occasionalism is the view espoused by Nicholas Malebranche as well as Islamic philosophers such as Abu Hamid Muhammad ibn Muhammad al-Ghazali that asserts that all supposedly causal relations between physical events, or between physical and mental events, are not really causal at all. While body and mind are different substances, causes (whether mental or physical) are related to their effects by an act of God's intervention on each specific occasion.[45]

Property dualism

Property dualism is the view that the world is constituted of just one kind of substance – the physical kind – and there exist two distinct kinds of properties: physical properties and mental properties. In other words, it is the view that non-physical, mental properties (such as beliefs, desires and emotions) inhere in some physical bodies (at least, brains). How mental and physical properties relate causally depends on the variety of property dualism in question, and is not always a clear issue. Sub-varieties of property dualism include:
  1. Strong emergentism asserts that when matter is organized in the appropriate way (i.e. in the way that living human bodies are organized), mental properties emerge in a way not fully accountable for by physical laws. Hence, it is a form of emergent materialism.[8] These emergent properties have an independent ontological status and cannot be reduced to, or explained in terms of, the physical substrate from which they emerge. They are dependent on the physical properties from which they emerge, but opinions vary as to the coherence of top–down causation, i.e. the causal effectiveness of such properties. A form of property dualism has been espoused by David Chalmers and the concept has undergone something of a renaissance in recent years,[46] but was already suggested in the 19th century by William James.
  2. Epiphenomenalism is a doctrine first formulated by Thomas Henry Huxley.[47] It consists of the view that mental phenomena are causally ineffectual, where one or more mental states do not have any influence on physical states or mental phenomena are the effects, but not the causes, of physical phenomena. Physical events can cause other physical events and physical events can cause mental events, but mental events cannot cause anything, since they are just causally inert by-products (i.e. epiphenomena) of the physical world.[43] This view has been defended most strongly in recent times by Frank Jackson.[48]
  3. Non-reductive Physicalism is the view that mental properties form a separate ontological class to physical properties: mental states (such as qualia) are not reducible to physical states. The ontological stance towards qualia in the case of non-reductive physicalism does not imply that qualia are causally inert; this is what distinguishes it from epiphenomenalism.
  4. Panpsychism is the view that all matter has a mental aspect, or, alternatively, all objects have a unified center of experience or point of view. Superficially, it seems to be a form of property dualism, since it regards everything as having both mental and physical properties. However, some panpsychists say mechanical behaviour is derived from primitive mentality of atoms and molecules—as are sophisticated mentality and organic behaviour, the difference being attributed to the presence or absence of complex structure in a compound object. So long as the reduction of non-mental properties to mental ones is in place, panpsychism is not a (strong) form of property dualism; otherwise it is.

Dual aspect theory

Dual aspect theory or dual-aspect monism is the view that the mental and the physical are two aspects of, or perspectives on, the same substance. (Thus it is a mixed position, which is monistic in some respects). In modern philosophical writings, the theory's relationship to neutral monism has become somewhat ill-defined, but one proffered distinction says that whereas neutral monism allows the context of a given group of neutral elements and the relationships into which they enter to determine whether the group can be thought of as mental, physical, both, or neither, dual-aspect theory suggests that the mental and the physical are manifestations (or aspects) of some underlying substance, entity or process that is itself neither mental nor physical as normally understood. Various formulations of dual-aspect monism also require the mental and the physical to be complementary, mutually irreducible and perhaps inseparable (though distinct).

Experiential dualism

This is a philosophy of mind that regards the degrees of freedom between mental and physical well-being as not necessarily synonymous thus implying an experiential dualism between body and mind. An example of these disparate degrees of freedom is given by Allan Wallace who notes that it is "experientially apparent that one may be physically uncomfortable—for instance, while engaging in a strenuous physical workout—while mentally cheerful; conversely, one may be mentally distraught while experiencing physical comfort".[52] Experiential dualism notes that our subjective experience of merely seeing something in the physical world seems qualitatively different than mental processes like grief that comes from losing a loved one. This philosophy also is a proponent of causal dualism which is defined as the dual ability for mental states and physical states to affect one another. Mental states can cause changes in physical states and vice versa.

However, unlike cartesian dualism or some other systems, experiential dualism does not posit two fundamental substances in reality: mind and matter. Rather, experiential dualism is to be understood as a conceptual framework that gives credence to the qualitative difference between the experience of mental and physical states. Experiential dualism is accepted as the conceptual framework of Madhyamaka Buddhism.

Madhayamaka Buddhism goes even further, finding fault with the monist view of physicalist philosophies of mind as well in that these generally posit matter and energy as the fundamental substance of reality. Nonetheless, this does not imply that the cartesian dualist view is correct, rather Madhyamaka regards as error any affirming view of a fundamental substance to reality.
In denying the independent self-existence of all the phenomena that make up the world of our experience, the Madhyamaka view departs from both the substance dualism of Descartes and the substance monism—namely, physicalism—that is characteristic of modern science. The physicalism propounded by many contemporary scientists seems to assert that the real world is composed of physical things-in-themselves, while all mental phenomena are regarded as mere appearances, devoid of any reality in and of themselves. Much is made of this difference between appearances and reality.[52]
Indeed, physicalism, or the idea that matter is the only fundamental substance of reality, is explicitly rejected by Buddhism.
In the Madhyamaka view, mental events are no more or less real than physical events. In terms of our common-sense experience, differences of kind do exist between physical and mental phenomena. While the former commonly have mass, location, velocity, shape, size, and numerous other physical attributes, these are not generally characteristic of mental phenomena. For example, we do not commonly conceive of the feeling of affection for another person as having mass or location. These physical attributes are no more appropriate to other mental events such as sadness, a recalled image from one's childhood, the visual perception of a rose, or consciousness of any sort. Mental phenomena are, therefore, not regarded as being physical, for the simple reason that they lack many of the attributes that are uniquely characteristic of physical phenomena. Thus, Buddhism has never adopted the physicalist principle that regards only physical things as real.[52]

Hylomorphism

Hylomorphism is a theory that originates with Aristotelian philosophy, which conceives being as a compound of matter and form. "Hylomorphism" is a 19th-century term formed from the Greek words ὕλη hyle, "wood, matter", and μορφή, morphē, "form".

Monist solutions to the mind–body problem

In contrast to dualism, monism does not accept any fundamental divisions. The fundamentally disparate nature of reality has been central to forms of eastern philosophies for over two millennia. In Indian and Chinese philosophy, monism is integral to how experience is understood. Today, the most common forms of monism in Western philosophy are physicalist.[18] Physicalistic monism asserts that the only existing substance is physical, in some sense of that term to be clarified by our best science.[53] However, a variety of formulations (see below) are possible. Another form of monism, idealism, states that the only existing substance is mental. Although pure idealism, such as that of George Berkeley, is uncommon in contemporary Western philosophy, a more sophisticated variant called panpsychism, according to which mental experience and properties may be at the foundation of physical experience and properties, has been espoused by some philosophers such as Alfred North Whitehead[54] and David Ray Griffin.[46]

Phenomenalism is the theory that representations (or sense data) of external objects are all that exist. Such a view was briefly adopted by Bertrand Russell and many of the logical positivists during the early 20th century.[55] A third possibility is to accept the existence of a basic substance that is neither physical nor mental. The mental and physical would then both be properties of this neutral substance. Such a position was adopted by Baruch Spinoza[9] and was popularized by Ernst Mach[56] in the 19th century. This neutral monism, as it is called, resembles property dualism.

Physicalistic monisms

Behaviorism

Behaviorism dominated philosophy of mind for much of the 20th century, especially the first half.[18] In psychology, behaviorism developed as a reaction to the inadequacies of introspectionism.[53] Introspective reports on one's own interior mental life are not subject to careful examination for accuracy and cannot be used to form predictive generalizations. Without generalizability and the possibility of third-person examination, the behaviorists argued, psychology cannot be scientific.[53] The way out, therefore, was to eliminate the idea of an interior mental life (and hence an ontologically independent mind) altogether and focus instead on the description of observable behavior.[57]
Parallel to these developments in psychology, a philosophical behaviorism (sometimes called logical behaviorism) was developed.[53] This is characterized by a strong verificationism, which generally considers unverifiable statements about interior mental life pointless. For the behaviorist, mental states are not interior states on which one can make introspective reports. They are just descriptions of behavior or dispositions to behave in certain ways, made by third parties to explain and predict another's behavior.[58]

Philosophical behaviorism has fallen out of favor since the latter half of the 20th century, coinciding with the rise of cognitivism.[2] Cognitivists reject behaviorism due to several perceived problems. For example, behaviorism could be said to be counterintuitive when it maintains that someone is talking about behavior in the event that a person is experiencing a painful headache.

Identity theory

Type physicalism (or type-identity theory) was developed by John Smart[25] and Ullin Place[59] as a direct reaction to the failure of behaviorism. These philosophers reasoned that, if mental states are something material, but not behavioral, then mental states are probably identical to internal states of the brain. In very simplified terms: a mental state M is nothing other than brain state B. The mental state "desire for a cup of coffee" would thus be nothing more than the "firing of certain neurons in certain brain regions".[25]
 
The classic Identity theory and Anomalous Monism in contrast. For the Identity theory, every token instantiation of a single mental type corresponds (as indicated by the arrows) to a physical token of a single physical type. For anomalous monism, the token–token correspondences can fall outside of the type–type correspondences. The result is token identity.

Despite its initial plausibility, the identity theory faces a strong challenge in the form of the thesis of multiple realizability, first formulated by Hilary Putnam.[27] It is obvious that not only humans, but many different species of animals can, for example, experience pain. However, it seems highly unlikely that all of these diverse organisms with the same pain experience are in the identical brain state. And if this is the case, then pain cannot be identical to a specific brain state. The identity theory is thus empirically unfounded.[27]

On the other hand, even granted the above, it does not follow that identity theories of all types must be abandoned. According to token identity theories, the fact that a certain brain state is connected with only one mental state of a person does not have to mean that there is an absolute correlation between types of mental state and types of brain state. The type–token distinction can be illustrated by a simple example: the word "green" contains four types of letters (g, r, e, n) with two tokens (occurrences) of the letter e along with one each of the others. The idea of token identity is that only particular occurrences of mental events are identical with particular occurrences or tokenings of physical events.[60] Anomalous monism (see below) and most other non-reductive physicalisms are token-identity theories.[61] Despite these problems, there is a renewed interest in the type identity theory today, primarily due to the influence of Jaegwon Kim.[25]

Functionalism

Functionalism was formulated by Hilary Putnam and Jerry Fodor as a reaction to the inadequacies of the identity theory.[27] Putnam and Fodor saw mental states in terms of an empirical computational theory of the mind.[62] At about the same time or slightly after, D.M. Armstrong and David Kellogg Lewis formulated a version of functionalism that analyzed the mental concepts of folk psychology in terms of functional roles.[63] Finally, Wittgenstein's idea of meaning as use led to a version of functionalism as a theory of meaning, further developed by Wilfrid Sellars and Gilbert Harman. Another one, psychofunctionalism, is an approach adopted by the naturalistic philosophy of mind associated with Jerry Fodor and Zenon Pylyshyn.
What all these different varieties of functionalism share in common is the thesis that mental states are characterized by their causal relations with other mental states and with sensory inputs and behavioral outputs. That is, functionalism abstracts away from the details of the physical implementation of a mental state by characterizing it in terms of non-mental functional properties. For example, a kidney is characterized scientifically by its functional role in filtering blood and maintaining certain chemical balances. From this point of view, it does not really matter whether the kidney be made up of organic tissue, plastic nanotubes or silicon chips: it is the role that it plays and its relations to other organs that define it as a kidney.[62]

Non-reductive physicalism

Non-reductionist philosophers hold firmly to two essential convictions with regard to mind–body relations: 1) Physicalism is true and mental states must be physical states, but 2) All reductionist proposals are unsatisfactory: mental states cannot be reduced to behavior, brain states or functional states.[53] Hence, the question arises whether there can still be a non-reductive physicalism. Donald Davidson's anomalous monism[26] is an attempt to formulate such a physicalism. He "thinks that when one runs across what are traditionally seen as absurdities of Reason, such as akrasia or self-deception, the personal psychology framework is not to be given up in favor of the subpersonal one, but rather must be enlarged or extended so that the rationality set out by the principle of charity can be found elsewhere."[64]
Davidson uses the thesis of supervenience: mental states supervene on physical states, but are not reducible to them. "Supervenience" therefore describes a functional dependence: there can be no change in the mental without some change in the physical–causal reducibility between the mental and physical without ontological reducibility.[65]

Because non-reductive physicalist theories attempt to both retain the ontological distinction between mind and body and to try to solve the "surfeit of explanations puzzle" in some way; critics often see this as a paradox and point out the similarities to epiphenomenalism, in that it is the brain that is seen as the root "cause" not the mind, and the mind seems to be rendered inert.

Epiphenomenalism regards one or more mental states as the byproduct of physical brain states, having no influence on physical states. The interaction is one-way (solving the "surfeit of explanations puzzle") but leaving us with non-reducible mental states (as a byproduct of brain states) – causally reducible, but ontologically irreducible to physical states. Pain would be seen by epiphenomenaliasts as being caused by the brain state but as not having effects on other brain states, though it might have effects on other mental states (i.e. cause distress).

Weak emergentism

Weak emergentism is a form of "non-reductive physicalism" that involves a layered view of nature, with the layers arranged in terms of increasing complexity and each corresponding to its own special science. Some philosophers hold that emergent properties causally interact with more fundamental levels, while others maintain that higher-order properties simply supervene over lower levels without direct causal interaction. The latter group therefore holds a less strict, or "weaker", definition of emergentism, which can be rigorously stated as follows: a property P of composite object O is emergent if it is metaphysically impossible for another object to lack property P if that object is composed of parts with intrinsic properties identical to those in O and has those parts in an identical configuration.
Sometimes emergentists use the example of water having a new property when Hydrogen H and Oxygen O combine to form H2O (water). In this example there "emerges" a new property of a transparent liquid that would not have been predicted by understanding hydrogen and oxygen as gases. This is analogous to physical properties of the brain giving rise to a mental state. Emergentists try to solve the notorious mind–body gap this way. One problem for emergentism is the idea of "causal closure" in the world that does not allow for a mind-to-body causation.[66]

Eliminative materialism

If one is a materialist and believes that all aspects of our common-sense psychology will find reduction to a mature cognitive neuroscience, and that non-reductive materialism is mistaken, then one can adopt a final, more radical position: eliminative materialism.
There are several varieties of eliminative materialism, but all maintain that our common-sense "folk psychology" badly misrepresents the nature of some aspect of cognition. Eliminativists such as Patricia and Paul Churchland argue that while folk psychology treats cognition as fundamentally sentence-like, the non-linguistic vector/matrix model of neural network theory or connectionism will prove to be a much more accurate account of how the brain works.[23]

The Churchlands often invoke the fate of other, erroneous popular theories and ontologies that have arisen in the course of history.[23][24] For example, Ptolemaic astronomy served to explain and roughly predict the motions of the planets for centuries, but eventually this model of the solar system was eliminated in favor of the Copernican model. The Churchlands believe the same eliminative fate awaits the "sentence-cruncher" model of the mind in which thought and behavior are the result of manipulating sentence-like states called "propositional attitudes".

Non-physicalist monisms

Idealism

Idealism is the form of monism that sees the world as consisting of minds, mental contents and or consciousness. Idealists are not faced with explaining how minds arise from bodies: rather, the world, bodies and objects are regarded as mere appearances held by minds. However, accounting for the mind–body problem is not usually the main motivation for idealism; rather, idealists tend to be motivated by skepticism, intentionality, and the unique nature of ideas. Idealism is prominent in Eastern religious and philosophical thought. It has gone through several cycles of popularity and neglect in the history of Western philosophy.

Different varieties of idealism may hold that there are

Neutral monism

Neutral monism, in philosophy, is the metaphysical view that the mental and the physical are two ways of organizing or describing the same elements, which are themselves "neutral", that is, neither physical nor mental. This view denies that the mental and the physical are two fundamentally different things. Rather, neutral monism claims the universe consists of only one kind of stuff, in the form of neutral elements that are in themselves neither mental nor physical. These neutral elements might have the properties of color and shape, just as we experience those properties. But these shaped and colored elements do not exist in a mind (considered as a substantial entity, whether dualistically or physicalistically); they exist on their own.

Mysterianism

Some philosophers take an epistemic approach and argue that the mind–body problem is currently unsolvable, and perhaps will always remain unsolvable to human beings. This is usually termed New mysterianism. Colin McGinn holds that human beings are cognitively closed in regards to their own minds. According to McGinn human minds lack the concept-forming procedures to fully grasp how mental properties such as consciousness arise from their causal basis.[67] An example would be how an elephant is cognitively closed in regards to particle physics.
A more moderate conception has been expounded by Thomas Nagel, which holds that the mind–body problem is currently unsolvable at the present stage of scientific development and that it might take a future scientific paradigm shift or revolution to bridge the explanatory gap. Nagel posits that in the future a sort of "objective phenomenology" might be able to bridge the gap between subjective conscious experience and its physical basis.[68]

Linguistic criticism of the mind–body problem

Each attempt to answer the mind–body problem encounters substantial problems. Some philosophers argue that this is because there is an underlying conceptual confusion.[69] These philosophers, such as Ludwig Wittgenstein and his followers in the tradition of linguistic criticism, therefore reject the problem as illusory.[70] They argue that it is an error to ask how mental and biological states fit together. Rather it should simply be accepted that human experience can be described in different ways—for instance, in a mental and in a biological vocabulary. Illusory problems arise if one tries to describe the one in terms of the other's vocabulary or if the mental vocabulary is used in the wrong contexts.[70] This is the case, for instance, if one searches for mental states of the brain. The brain is simply the wrong context for the use of mental vocabulary—the search for mental states of the brain is therefore a category error or a sort of fallacy of reasoning.[70]

Today, such a position is often adopted by interpreters of Wittgenstein such as Peter Hacker.[69] However, Hilary Putnam, the originator of functionalism, has also adopted the position that the mind–body problem is an illusory problem which should be dissolved according to the manner of Wittgenstein.[71]

Externalism and internalism

Where is the mind located? If the mind is a physical phenomenon of some kind, it has to be located somewhere. According to some, there are two possible options: either the mind is internal to the body (internalism) or the mind is external to it (externalism). More generally, either the mind depends only on events and properties taking place inside the subject's body or it depends also on factors external to it.

Proponents of internalism are committed to the view that neural activity is sufficient to produce the mind. Proponents of externalism maintain that the surrounding world is in some sense constitutive of the mind.

Externalism differentiates into several versions. The main ones are semantic externalism, cognitive externalism and phenomenal externalism. Each of these versions of externalism can further be divided into whether they refer only to the content or to the vehicles of the mind.

Semantic externalism holds that the semantic content of the mind is totally or partially defined by a state of affairs external to the body of the subject. Hilary Putnam's Twin Earth thought experiment is a good example.

Cognitive externalism is a very broad collection of views that suggests the role of the environment, of tools, of development, and of the body in fleshing out cognition. Embodied cognition, the extended mind, and enactivism are good examples.

Phenomenal externalism suggests that the phenomenal aspects of the mind are external to the body. Authors who addressed this possibility are Ted Honderich, Edwin Holt, Francois Tonneau, Kevin O'Regan, Riccardo Manzotti, Teed Rockwell and Max Velmans.

Naturalism and its problems

The thesis of physicalism is that the mind is part of the material (or physical) world. Such a position faces the problem that the mind has certain properties that no other material thing seems to possess. Physicalism must therefore explain how it is possible that these properties can nonetheless emerge from a material thing. The project of providing such an explanation is often referred to as the "naturalization of the mental".[53] Some of the crucial problems that this project attempts to resolve include the existence of qualia and the nature of intentionality.[53]

Qualia

Many mental states seem to be experienced subjectively in different ways by different individuals.[32] And it is characteristic of a mental state that it has some experiential quality, e.g. of pain, that it hurts. However, the sensation of pain between two individuals may not be identical, since no one has a perfect way to measure how much something hurts or of describing exactly how it feels to hurt. Philosophers and scientists therefore ask where these experiences come from. The existence of cerebral events, in and of themselves, cannot explain why they are accompanied by these corresponding qualitative experiences. The puzzle of why many cerebral processes occur with an accompanying experiential aspect in consciousness seems impossible to explain.[31]

Yet it also seems to many that science will eventually have to explain such experiences.[53] This follows from an assumption about the possibility of reductive explanations. According to this view, if an attempt can be successfully made to explain a phenomenon reductively (e.g., water), then it can be explained why the phenomenon has all of its properties (e.g., fluidity, transparency).[53] In the case of mental states, this means that there needs to be an explanation of why they have the property of being experienced in a certain way.

The 20th-century German philosopher Martin Heidegger criticized the ontological assumptions underpinning such a reductive model, and claimed that it was impossible to make sense of experience in these terms. This is because, according to Heidegger, the nature of our subjective experience and its qualities is impossible to understand in terms of Cartesian "substances" that bear "properties". Another way to put this is that the very concept of qualitative experience is incoherent in terms of—or is semantically incommensurable with the concept of—substances that bear properties.[72]

This problem of explaining introspective first-person aspects of mental states and consciousness in general in terms of third-person quantitative neuroscience is called the explanatory gap.[73] There are several different views of the nature of this gap among contemporary philosophers of mind. David Chalmers and the early Frank Jackson interpret the gap as ontological in nature; that is, they maintain that qualia can never be explained by science because physicalism is false. There are two separate categories involved and one cannot be reduced to the other.[74] An alternative view is taken by philosophers such as Thomas Nagel and Colin McGinn. According to them, the gap is epistemological in nature. For Nagel, science is not yet able to explain subjective experience because it has not yet arrived at the level or kind of knowledge that is required. We are not even able to formulate the problem coherently.[32] For McGinn, on other hand, the problem is one of permanent and inherent biological limitations. We are not able to resolve the explanatory gap because the realm of subjective experiences is cognitively closed to us in the same manner that quantum physics is cognitively closed to elephants.[75] Other philosophers liquidate the gap as purely a semantic problem. This semantic problem, of course, led to the famous "Qualia Question", which is: Does Red cause Redness?

Intentionality

John Searle—one of the most influential philosophers of mind, proponent of biological naturalism (Berkeley 2002)

Intentionality is the capacity of mental states to be directed towards (about) or be in relation with something in the external world.[29] This property of mental states entails that they have contents and semantic referents and can therefore be assigned truth values. When one tries to reduce these states to natural processes there arises a problem: natural processes are not true or false, they simply happen.[76] It would not make any sense to say that a natural process is true or false. But mental ideas or judgments are true or false, so how then can mental states (ideas or judgments) be natural processes? The possibility of assigning semantic value to ideas must mean that such ideas are about facts. Thus, for example, the idea that Herodotus was a historian refers to Herodotus and to the fact that he was a historian. If the fact is true, then the idea is true; otherwise, it is false. But where does this relation come from? In the brain, there are only electrochemical processes and these seem not to have anything to do with Herodotus.[28]

Philosophy of perception

Philosophy of perception is concerned with the nature of perceptual experience and the status of perceptual objects, in particular how perceptual experience relates to appearances and beliefs about the world. The main contemporary views within philosophy of perception include naive realism, enactivism and representational views.[3][4][77]

Philosophy of mind and science

Humans are corporeal beings and, as such, they are subject to examination and description by the natural sciences. Since mental processes are intimately related to bodily processes, the descriptions that the natural sciences furnish of human beings play an important role in the philosophy of mind.[2] There are many scientific disciplines that study processes related to the mental. The list of such sciences includes: biology, computer science, cognitive science, cybernetics, linguistics, medicine, pharmacology, and psychology.[78]

Neurobiology

The theoretical background of biology, as is the case with modern natural sciences in general, is fundamentally materialistic. The objects of study are, in the first place, physical processes, which are considered to be the foundations of mental activity and behavior.[79] The increasing success of biology in the explanation of mental phenomena can be seen by the absence of any empirical refutation of its fundamental presupposition: "there can be no change in the mental states of a person without a change in brain states."[78]
Within the field of neurobiology, there are many subdisciplines that are concerned with the relations between mental and physical states and processes:[79] Sensory neurophysiology investigates the relation between the processes of perception and stimulation.[80] Cognitive neuroscience studies the correlations between mental processes and neural processes.[80] Neuropsychology describes the dependence of mental faculties on specific anatomical regions of the brain.[80] Lastly, evolutionary biology studies the origins and development of the human nervous system and, in as much as this is the basis of the mind, also describes the ontogenetic and phylogenetic development of mental phenomena beginning from their most primitive stages.[78] Evolutionary biology furthermore places tight constraints on any philosophical theory of the mind, as the gene-based mechanism of natural selection does not allow any giant leaps in the development of neural complexity or neural software but only incremental steps over long time periods.[81]
 
Since the 1980s, sophisticated neuroimaging procedures, such as fMRI (above), have furnished increasing knowledge about the workings of the human brain, shedding light on ancient philosophical problems.

The methodological breakthroughs of the neurosciences, in particular the introduction of high-tech neuroimaging procedures, has propelled scientists toward the elaboration of increasingly ambitious research programs: one of the main goals is to describe and comprehend the neural processes which correspond to mental functions (see: neural correlate).[79] Several groups are inspired by these advances.

Computer science

Computer science concerns itself with the automatic processing of information (or at least with physical systems of symbols to which information is assigned) by means of such things as computers.[82] From the beginning, computer programmers have been able to develop programs that permit computers to carry out tasks for which organic beings need a mind. A simple example is multiplication. It is not clear whether computers could be said to have a mind. Could they, someday, come to have what we call a mind? This question has been propelled into the forefront of much philosophical debate because of investigations in the field of artificial intelligence (AI).
Within AI, it is common to distinguish between a modest research program and a more ambitious one: this distinction was coined by John Searle in terms of a weak AI and strong AI. The exclusive objective of "weak AI", according to Searle, is the successful simulation of mental states, with no attempt to make computers become conscious or aware, etc. The objective of strong AI, on the contrary, is a computer with consciousness similar to that of human beings.[83] The program of strong AI goes back to one of the pioneers of computation Alan Turing. As an answer to the question "Can computers think?", he formulated the famous Turing test.[84] Turing believed that a computer could be said to "think" when, if placed in a room by itself next to another room that contained a human being and with the same questions being asked of both the computer and the human being by a third party human being, the computer's responses turned out to be indistinguishable from those of the human. Essentially, Turing's view of machine intelligence followed the behaviourist model of the mind—intelligence is as intelligence does. The Turing test has received many criticisms, among which the most famous is probably the Chinese room thought experiment formulated by Searle.[83]

The question about the possible sensitivity (qualia) of computers or robots still remains open. Some computer scientists believe that the specialty of AI can still make new contributions to the resolution of the "mind–body problem". They suggest that based on the reciprocal influences between software and hardware that takes place in all computers, it is possible that someday theories can be discovered that help us to understand the reciprocal influences between the human mind and the brain (wetware).[85]

Psychology

Psychology is the science that investigates mental states directly. It uses generally empirical methods to investigate concrete mental states like joy, fear or obsessions. Psychology investigates the laws that bind these mental states to each other or with inputs and outputs to the human organism.[86]
An example of this is the psychology of perception. Scientists working in this field have discovered general principles of the perception of forms. A law of the psychology of forms says that objects that move in the same direction are perceived as related to each other.[78] This law describes a relation between visual input and mental perceptual states. However, it does not suggest anything about the nature of perceptual states. The laws discovered by psychology are compatible with all the answers to the mind–body problem already described.

Cognitive science

Cognitive science is the interdisciplinary scientific study of the mind and its processes. It examines what cognition is, what it does, and how it works. It includes research on intelligence and behavior, especially focusing on how information is represented, processed, and transformed (in faculties such as perception, language, memory, reasoning, and emotion) within nervous systems (human or other animal) and machines (e.g. computers). Cognitive science consists of multiple research disciplines, including psychology, artificial intelligence, philosophy, neuroscience, linguistics, anthropology, sociology, and education.[87] It spans many levels of analysis, from low-level learning and decision mechanisms to high-level logic and planning; from neural circuitry to modular brain organisation. Rowlands argues that cognition is enactive, embodied, embedded, affective and (potentially) extended. The position is taken that the "classical sandwich" of cognition sandwiched between perception and action is artificial; cognition has to be seen as a product of a strongly coupled interaction that cannot be divided this way.[88][89]

Philosophy of mind in the continental tradition

Most of the discussion in this article has focused on one style or tradition of philosophy in modern Western culture, usually called analytic philosophy (sometimes described as Anglo-American philosophy).[90] Many other schools of thought exist, however, which are sometimes subsumed under the broad (and vague) label of continental philosophy.[90] In any case, though topics and methods here are numerous, in relation to the philosophy of mind the various schools that fall under this label (phenomenology, existentialism, etc.) can globally be seen to differ from the analytic school in that they focus less on language and logical analysis alone but also take in other forms of understanding human existence and experience. With reference specifically to the discussion of the mind, this tends to translate into attempts to grasp the concepts of thought and perceptual experience in some sense that does not merely involve the analysis of linguistic forms.[90]

Immanuel Kant's Critique of Pure Reason, first published in 1781 and presented again with major revisions in 1787, represents a significant intervention into what will later become known as the philosophy of mind. Kant's first critique is generally recognized as among the most significant works of modern philosophy in the West. Kant is a figure whose influence is marked in both continental and analytic/Anglo-American philosophy. Kant's work develops an in-depth study of transcendental consciousness, or the life of the mind as conceived through universal categories of consciousness.

In Georg Wilhelm Friedrich Hegel's Philosophy of Mind (frequently translated as Philosophy of Spirit or Geist),[91] the third part of his Encyclopedia of the Philosophical Sciences, Hegel discusses three distinct types of mind: the "subjective mind/spirit", the mind of an individual; the "objective mind/spirit", the mind of society and of the State; and the "Absolute mind/spirit", the position of religion, art, and philosophy. See also Hegel's The Phenomenology of Spirit. Nonetheless, Hegel's work differs radically from the style of Anglo-American philosophy of mind.

In 1896, Henri Bergson made in Matter and Memory "Essay on the relation of body and spirit" a forceful case for the ontological difference of body and mind by reducing the problem to the more definite one of memory, thus allowing for a solution built on the empirical test case of aphasia.

In modern times, the two main schools that have developed in response or opposition to this Hegelian tradition are phenomenology and existentialism. Phenomenology, founded by Edmund Husserl, focuses on the contents of the human mind (see noema) and how processes shape our experiences.[92] Existentialism, a school of thought founded upon the work of Søren Kierkegaard, focuses on Human predicament and how people deal with the situation of being alive. Existential-phenomenology represents a major branch of continental philosophy (they are not contradictory), rooted in the work of Husserl but expressed in its fullest forms in the work of Martin Heidegger, Jean-Paul Sartre, Simone de Beauvoir and Maurice Merleau-Ponty. See Heidegger's Being and Time, Merleau-Ponty's Phenomenology of Perception, Sartre's Being and Nothingness, and Simone de Beauvoir's The Second Sex.

Mind in Eastern philosophy

Mind in Hindu philosophy

Dualism

Substance Dualism is a common feature of several orthodox Hindu schools including the Sāṅkhya, Nyāya, Yoga and Dvaita Vedanta. In these schools a clear difference is drawn between matter and a non-material soul, which is eternal and undergoes samsara, a cycle of death and rebirth. The Nyāya school argued that qualities such as cognition and desire are inherent qualities which are not possessed by anything solely material, and therefore by process of elimination must belong to a non-material self, the atman.[93] Many of these schools see their spiritual goal as moksha, liberation from the cycle of reincarnation.

Vedanta monistic idealism

Śaṅkara

In the Advaita Vedanta of the 8th century Indian philosopher Śaṅkara, the mind, body and world are all held to be the same unchanging eternal conscious entity called Brahman. Advaita, which means non-dualism, holds the view that all that exists is pure absolute consciousness. The fact that the world seems to be made up of changing entities is an illusion, or Maya. The only thing that exists is Brahman, which is described as Satchitananda (Being, consciousness and bliss). Advaita Vedanta is best described by a verse which states "Brahman is alone True, and this world of plurality is an error; the individual self is not different from Brahman."[94]

Another form of monistic Vedanta is Vishishtadvaita (qualified non-dualism) as posited by the eleventh century philosopher Ramanuja. Ramanuja criticized Advaita Vedanta by arguing that consciousness is always intentional and that it is also always a property of something. Ramanuja's Brahman is defined by a multiplicity of qualities and properties in a single monistic entity. This doctrine is called "samanadhikaranya" (several things in a common substrate).[95]

Materialism

Arguably the first exposition of empirical materialism in the history of philosophy is in the Cārvāka school (also called Lokāyata). The Cārvāka school rejected the existence of anything but matter (which they defined as being made up of the four elements), including God and the soul. Therefore, they held that even consciousness was nothing but a construct made up of atoms. A section of the Cārvāka school believed in a material soul made up of air or breath, but since this also was a form of matter, it was not said to survive death.[96]

Buddhist philosophy of mind

 The Five Aggregates (pañca khandha)
according to the Pali Canon.
 
 
form (rūpa)
  4 elements
(mahābhūta)
 
 
   
    contact
(phassa)
    

consciousness
(viññāna)

 









  mental factors (cetasika)  

feeling
(vedanā)

 
 

perception
(sañña)

 
 

formation
(saṅkhāra)

 
 
 
 
 Source: MN 109 (Thanissaro, 2001)  |  diagram details
Buddhist teachings describe that the mind manifests moment-to-moment as sense impressions and mental phenomena that are continuously changing.[97] The moment-by-moment manifestation of the mind-stream has been described as happening in every person all the time, even in a scientist who analyses various phenomena in the world, or analyses the material body including the organ brain.[97] The manifestation of the mind-stream is also described as being influenced by physical laws, biological laws, psychological laws, volitional laws, and universal laws.[97]

A salient feature of Buddhist philosophy which sets it apart from Indian orthodoxy is the centrality of the doctrine of not-self (Pāli. anatta, Skt. anātman). The Buddha's not-self doctrine sees humans as an impermanent composite of five psychological and physical aspects instead of a single fixed self. In this sense, what is called ego or the self is merely a convenient fiction, an illusion that does not apply to anything real but to an erroneous way of looking at the ever-changing stream of five interconnected aggregate factors.[98] The relationship between these aggregates is said to be one of dependent-arising (pratītyasamutpāda).
This means that all things, including mental events, arise co-dependently from a plurality of other causes and conditions. This seems to reject both causal determinist and epiphenomenalist conceptions of mind.[98]

Abhidharma theories of mind

Three centuries after the death of the Buddha (c. 150 BCE) saw the growth of a large body of literature called the Abhidharma in several contending Buddhist schools. In the Abhidharmic analysis of mind, the ordinary thought is defined as prapañca ('conceptual proliferation'). According to this theory, perceptual experience is bound up in multiple conceptualizations (expectations, judgments and desires). This proliferation of conceptualizations form our illusory superimposition of concepts like self and other upon an ever-changing stream of aggregate phenomena.[98] In this conception of mind no strict distinction is made between the conscious faculty and the actual sense perception of various phenomena. Consciousness is instead said to be divided into six sense modalities, five for the five senses and sixth for perception of mental phenomena.[98] The arising of cognitive awareness is said to depend on sense perception, awareness of the mental faculty itself which is termed mental or 'introspective awareness' (manovijñāna) and attention (āvartana), the picking out of objects out of the constantly changing stream of sensory impressions.

Rejection of a permanent agent eventually led to the philosophical problems of the seeming continuity of mind and also of explaining how rebirth and karma continue to be relevant doctrines without an eternal mind. This challenge was met by the Theravāda school by introducing the concept of mind as a factor of existence. This "life-stream" (Bhavanga-sota) is an undercurrent forming the condition of being. The continuity of a karmic "person" is therefore assured in the form of a mindstream (citta-santana), a series of flowing mental moments arising from the subliminal life-continuum mind (Bhavanga-citta), mental content, and attention.[98]

Indian Mahayana

The Sautrāntika school held a form of phenomenalism that saw the world as imperceptible. It held that external objects exist only as a support for cognition, which can only apprehend mental representations. This influenced the later Yogācāra school of Mahayana Buddhism. The Yogācāra school is often called the mind-only school because of its internalist stance that consciousness is the ultimate existing reality. The works of Vasubandhu have often been interpreted as arguing for some form of Idealism. Vasubandhu uses the dream argument and a mereological refutation of atomism to attack the reality of external objects as anything other than mental entities.[99] Scholarly interpretations of Vasubandhu's philosophy vary widely, and include phenomenalism, neutral monism and realist phenomenology.

The Indian Mahayana schools were divided on the issue of the possibility of reflexive awareness (svasaṃvedana). Dharmakīrti accepted the idea of reflexive awareness as expounded by the Yogācāra school, comparing it to a lamp that illuminates itself while also illuminating other objects. This was strictly rejected by Mādhyamika scholars like Candrakīrti. Since in the philosophy of the Mādhyamika all things and mental events are characterized by emptiness, they argued that consciousness could not be an inherently reflexive ultimate reality since that would mean it was self-validating and therefore not characterized by emptiness.[98] These views were ultimately reconciled by the 8th century thinker Śāntarakṣita. In Śāntarakṣita's synthesis he adopts the idealist Yogācāra views of reflexive awareness as a conventional truth into the structure of the two truths doctrine. Thus he states: "By relying on the Mind-Only system, know that external entities do not exist. And by relying on this Middle Way system, know that no self exists at all, even in that [mind]." [100]

The Yogācāra school also developed the theory of the repository consciousness (ālayavijñāna) to explain continuity of mind in rebirth and accumulation of karma. This repository consciousness acts as a storehouse for karmic seeds (bija) when all other senses are absent during the process of death and rebirth as well as being the causal potentiality of dharmic phenomena.[98] Thus according to B. Alan Wallace:
No constituents of the body—in the brain or elsewhere—transform into mental states and processes. Such subjective experiences do not emerge from the body, but neither do they emerge from nothing. Rather, all objective mental appearances arise from the substrate, and all subjective mental states and processes arise from the substrate consciousness.[101]

Tibetan Buddhism

Tibetan Buddhist theories of mind evolved directly from the Indian Mahayana views. Thus the founder of the Gelug school, Je Tsongkhapa discusses the Yogācāra system of the Eight Consciousnesses in his Explanation of the Difficult Points.[102] He would later come to repudiate Śāntarakṣita's pragmatic idealism. According to the 14th Dalai Lama the mind can be defined "as an entity that has the nature of mere experience, that is, 'clarity and knowing'. It is the knowing nature, or agency, that is called mind, and this is non-material."[103] The simultaneously dual nature of mind is as follows:
1. Clarity (gsal) – The mental activity which produces cognitive phenomena (snang-ba).
2. Knowing (rig) – The mental activity of perceiving cognitive phenomena.
The 14th Dalai Lama has also explicitly laid out his theory of mind as experiential dualism which is described above under the different types of dualism.[52]

Because Tibetan philosophy of mind is ultimately soteriological, it focuses on meditative practices such as Dzogchen and Mahamudra that allow a practitioner to experience the true reflexive nature of their mind directly. This unobstructed knowledge of one's primordial, empty and non-dual Buddha nature is called rigpa. The mind's innermost nature is described among various schools as pure luminosity or "clear light" ('od gsal) and is often compared to a crystal ball or a mirror. Sogyal Rinpoche speaks of mind thus: "Imagine a sky, empty, spacious, and pure from the beginning; its essence is like this. Imagine a sun, luminous, clear, unobstructed, and spontaneously present; its nature is like this."

Zen Buddhism

The central issue in Chinese Zen philosophy of mind is in the difference between the pure and awakened mind and the defiled mind. Chinese Chan master Huangpo described the mind as without beginning and without form or limit while the defiled mind was that which was obscured by attachment to form and concepts.[104] The pure Buddha-mind is thus able to see things "as they truly are", as absolute and non-dual "thusness" (Tathatā). This non-conceptual seeing also includes the paradoxical fact that there is no difference between a defiled and a pure mind, as well as no difference between samsara and nirvana.[104]

In the Shobogenzo, the Japanese philosopher Dogen argued that body and mind are neither ontologically nor phenomenologically distinct but are characterized by a oneness called shin jin (bodymind). According to Dogen, "casting off body and mind" (Shinjin datsuraku) in zazen will allow one to experience things-as-they-are (genjokoan) which is the nature of original enlightenment (hongaku).[105]

Topics related to philosophy of mind

There are countless subjects that are affected by the ideas developed in the philosophy of mind. Clear examples of this are the nature of death and its definitive character, the nature of emotion, of perception and of memory. Questions about what a person is and what his or her identity consists of also have much to do with the philosophy of mind. There are two subjects that, in connection with the philosophy of the mind, have aroused special attention: free will and the self.[2]

Free will

In the context of philosophy of mind, the problem of free will takes on renewed intensity. This is certainly the case, at least, for materialistic determinists.[2] According to this position, natural laws completely determine the course of the material world. Mental states, and therefore the will as well, would be material states, which means human behavior and decisions would be completely determined by natural laws. Some take this reasoning a step further: people cannot determine by themselves what they want and what they do. Consequently, they are not free.[106]
This argumentation is rejected, on the one hand, by the compatibilists. Those who adopt this position suggest that the question "Are we free?" can only be answered once we have determined what the term "free" means. The opposite of "free" is not "caused" but "compelled" or "coerced". It is not appropriate to identify freedom with indetermination. A free act is one where the agent could have done otherwise if it had chosen otherwise. In this sense a person can be free even though determinism is true.[106] The most important compatibilist in the history of the philosophy was David Hume.[107] More recently, this position is defended, for example, by Daniel Dennett.[108]

On the other hand, there are also many incompatibilists who reject the argument because they believe that the will is free in a stronger sense called libertarianism.[106] These philosophers affirm the course of the world is either a) not completely determined by natural law where natural law is intercepted by physically independent agency,[109] b) determined by indeterministic natural law only, or c) determined by indeterministic natural law in line with the subjective effort of physically non-reducible agency.[110] Under Libertarianism, the will does not have to be deterministic and, therefore, it is potentially free. Critics of the second proposition (b) accuse the incompatibilists of using an incoherent concept of freedom. They argue as follows: if our will is not determined by anything, then we desire what we desire by pure chance. And if what we desire is purely accidental, we are not free. So if our will is not determined by anything, we are not free.[106]

Self

The philosophy of mind also has important consequences for the concept of "self". If by "self" or "I" one refers to an essential, immutable nucleus of the person, some modern philosophers of mind, such as Daniel Dennett believe that no such thing exists. According to Dennett and other contemporaries, the self is considered an illusion.[111] The idea of a self as an immutable essential nucleus derives from the idea of an immaterial soul. Such an idea is unacceptable to modern philosophers with physicalist orientations and their general skepticism of the concept of "self" as postulated by David Hume, who could never catch himself not doing, thinking or feeling anything.[112] However, in the light of empirical results from developmental psychology, developmental biology and neuroscience, the idea of an essential inconstant, material nucleus—an integrated representational system distributed over changing patterns of synaptic connections—seems reasonable.

Black Wednesday

From Wikipedia, the free encyclopedia ...