an engineer wearing a helmet of sensors, part of a brain scanner.

Human memory: How we make, remember, and forget memories

Human memory happens in many parts of the brain at once, and some types of memories stick around longer than others.

From the moment we are born, our brains are bombarded by an immense amount of information about ourselves and the world around us. So, how do we hold on to everything we've learned and experienced? Memories.

Humans retain different types of memories for different lengths of time . Short-term memories last seconds to hours, while long-term memories last for years. We also have a working memory, which lets us keep something in our minds for a limited time by repeating it. Whenever you say a phone number to yourself over and over to remember it, you're using your working memory.

Another way to categorize memories is by the subject of the memory itself, and whether you are consciously aware of it. Declarative memory, also called explicit memory, consists of the sorts of memories you experience consciously. Some of these memories are facts or “common knowledge”: things like the capital of Portugal (Lisbon), or the number of cards in a standard deck of playing cards (52). Others consist of past events you've experienced, such as a childhood birthday.

Nondeclarative memory, also called implicit memory, unconsciously builds up. These include procedural memories, which your body uses to remember the skills you've learned. Do you play an instrument or ride a bicycle? Those are your procedural memories at work. Nondeclarative memories also can shape your body's unthinking responses, like salivating at the sight of your favorite food or tensing up when you see something you fear.

In general, declarative memories are easier to form than nondeclarative memories. It takes less time to memorize a country's capital than it does to learn how to play the violin. But nondeclarative memories stick around more easily. Once you've learned to ride a bicycle, you're not likely to forget.

The types of amnesia

To understand how we remember things, it's incredibly helpful to study how we forget— which is why neuroscientists study amnesia, the loss of memories or the ability to learn . Amnesia is usually the result of some kind of trauma to the brain, such as a head injury, a stroke, a brain tumor, or chronic alcoholism.

For Hungry Minds

There are two main types of amnesia. The first, retrograde amnesia, occurs where you forget things you knew before the brain trauma. Anterograde amnesia is when brain trauma curtails or stops someone's ability to form new memories.

The most famous case study of anterograde amnesia is Henry Molaison , who in 1953 had parts of his brain removed as a last-ditch treatment for severe seizures. While Molaison—known when he was alive as H.M.—remembered much of his childhood, he was unable to form new declarative memories. People who worked with him for decades had to re-introduce themselves with every visit.

You May Also Like

presentation on human memory

This neuroscientist makes memories—and suppresses them

presentation on human memory

She had a severe brain injury—so her husband turned to an unprecedented therapy

presentation on human memory

7 medical breakthroughs that gave us hope in 2023

By studying people such as H.M., as well as animals with different types of brain damage, scientists can trace where and how different kinds of memories form in the brain. It seems that short-term and long-term memories don't form in exactly the same way, nor do declarative and procedural memories.

There's no one place within the brain that holds all of your memories; different areas of the brain form and store different kinds of memories, and different processes may be at play for each. For instance, emotional responses such as fear reside in a brain region called the amygdala. Memories of the skills you've learned are associated with a different region called the striatum. A region called the hippocampus is crucial for forming, retaining, and recalling declarative memories. The temporal lobes, the brain regions that H.M. was partially missing, play a crucial role in forming and recalling memories.

How memories are formed, stored, and recalled

Since the 1940s scientists have surmised that memories are held within groups of neurons, or nerve cells, called cell assemblies. Those interconnected cells fire as a group in response to a specific stimulus, whether it's your friend's face or the smell of freshly baked bread. The more the neurons fire together, the more the cells' interconnections strengthen . That way, when a future stimulus triggers the cells, it's more likely that the whole assembly fires. The nerves' collective activity transcribes what we experience as a memory. Scientists are still working through the details of how it works.

For a short-term memory to become a long-term memory, it must be strengthened for long-term storage, a process called memory consolidation. Consolidation is thought to take place by several processes. One, called long-term potentiation, consists of individual nerves modifying themselves to grow and talk to their neighboring nerves differently. That remodeling alters the nerves' connections in the long term, which stabilizes the memory. All animals that have long-term memories use this same basic cellular machinery; scientists worked out the details of long-term potentiation by studying California sea slugs . However, not all long-term memories necessarily have to start as short-term memories.

As we recall a memory, many parts of our brain rapidly talk to each other, including regions in the brain's cortex that do high-level information processing, regions that handle our senses' raw inputs, and a region called the medial temporal lobe that seems to help coordinate the process. One recent study found that at the moment when patients recalled newly formed memories, ripples of nerve activity in the medial temporal lobe synced up with ripples in the brain's cortex.

Many mysteries of memory remain. How precisely are memories encoded within groups of neurons? How widely distributed in the brain are the cells that encode a given memory? How does our brain activity correspond to how we experience memories? These active areas of research may one day provide new insight into brain function and how to treat memory-related conditions .

For instance, recent work has demonstrated that some memories must be “reconsolidated” each time they're recalled. If so, the act of remembering something makes that memory temporarily malleable—letting it be strengthened, weakened, or otherwise altered. Memories may be more easily targeted by medications during reconsolidation, which could help treat conditions such as post-traumatic stress disorder, or PTSD .

Related Topics

  • NEUROSCIENCE

presentation on human memory

One way to heal a brain injury? Let LSD open your mind—literally.

presentation on human memory

What's the source of your headaches? It could be your neck.

presentation on human memory

She shined a light on the disease no one wants to see

presentation on human memory

Some people don’t experience stress. Are they happier?

presentation on human memory

What are the early warning signs of CTE? New clues emerge

  • Environment
  • Paid Content
  • Photography
  • Perpetual Planet

History & Culture

  • History & Culture
  • History Magazine
  • Mind, Body, Wonder
  • Terms of Use
  • Privacy Policy
  • Your US State Privacy Rights
  • Children's Online Privacy Policy
  • Interest-Based Ads
  • About Nielsen Measurement
  • Do Not Sell or Share My Personal Information
  • Nat Geo Home
  • Attend a Live Event
  • Book a Trip
  • Inspire Your Kids
  • Shop Nat Geo
  • Visit the D.C. Museum
  • Learn About Our Impact
  • Support Our Mission
  • Advertise With Us
  • Customer Service
  • Renew Subscription
  • Manage Your Subscription
  • Work at Nat Geo
  • Sign Up for Our Newsletters
  • Contribute to Protect the Planet

Copyright © 1996-2015 National Geographic Society Copyright © 2015-2024 National Geographic Partners, LLC. All rights reserved

Logo for Digital Editions

Want to create or adapt books like this? Learn more about how Pressbooks supports open publishing practices.

48 Memory and the Brain

Learning Objectives

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

  • Explain the brain functions involved in memory
  • Recognize the roles of the hippocampus, amygdala, and cerebellum

Are memories stored in just one part of the brain, or are they stored in many different parts of the brain? Karl Lashley began exploring this problem, about 100 years ago, by making lesions in the brains of animals such as rats and monkeys. He was searching for evidence of the  engram : the group of neurones that serve as the “physical representation of memory” (Josselyn, 2010). First, Lashley (1950) trained rats to find their way through a maze. Then, he used the tools available at the time—in this case a soldering iron—to create lesions in the rats’ brains, specifically in the cerebral cortex. He did this because he was trying to erase the engram, or the original memory trace that the rats had of the maze.

Lashley did not find evidence of the engram, and the rats were still able to find their way through the maze, regardless of the size or location of the lesion. Based on his creation of lesions and the animals’ reaction, he formulated the  equipotentiality hypothesis : if part of one area of the brain involved in memory is damaged, another part of the same area can take over that memory function (Lashley, 1950). Although Lashley’s early work did not confirm the existence of the engram, modern psychologists are making progress locating it. For example, Eric Kandel has spent decades studying the synapse and its role in controlling the flow of information through neural circuits needed to store memories (Mayford, Siegelbaum, & Kandel, 2012).

Many scientists believe that the entire brain is involved with memory. However, since Lashley’s research, other scientists have been able to look more closely at the brain and memory. They have argued that memory is located in specific parts of the brain, and specific neurones can be recognized for their involvement in forming memories. The main parts of the brain involved with memory are the amygdala, the hippocampus, the cerebellum, and the prefrontal cortex ( Figure M.20 ).

An illustration of a brain shows the location of the amygdala, hippocampus, cerebellum, and prefrontal cortex.

The Amygdala

First, let’s look at the role of the  amygdala  in memory formation. The main job of the amygdala is to regulate emotions, such as fear and aggression ( Figure M.20 ). The amygdala plays a part in how memories are stored because storage is influenced by stress hormones. For example, one researcher experimented with rats and the fear response (Josselyn, 2010). Using Pavlovian conditioning, a neutral tone was paired with a foot shock to the rats. This produced a fear memory in the rats. After being conditioned, each time they heard the tone, they would freeze (a defence response in rats), indicating a memory for the impending shock. Then the researchers induced cell death in neurones in the lateral amygdala, which is the specific area of the brain responsible for fear memories. They found the fear memory faded (became extinct). Because of its role in processing emotional information, the amygdala is also involved in memory consolidation: the process of transferring new learning into long-term memory. The amygdala seems to facilitate encoding memories at a deeper level when the event is emotionally arousing.

LINK TO LEARNING

The Hippocampus

Another group of researchers also experimented with rats to learn how the  hippocampus  functions in memory processing ( Figure M.20 ). They created lesions in the hippocampi of the rats, and found that the rats demonstrated memory impairment on various tasks, such as object recognition and maze running. They concluded that the hippocampus is involved in memory, specifically normal recognition memory as well as spatial memory (when the memory tasks are like recall tests) (Clark, Zola, & Squire, 2000). Another job of the hippocampus is to project information to cortical regions that give memories meaning and connect them with other memories. It also plays a part in memory consolidation: the process of transferring new learning into long-term memory.

Injury to this area leaves us unable to process new declarative memories. One famous patient, known for years only as H. M., had both his left and right temporal lobes (hippocampi) removed in an attempt to help control the seizures he had been suffering from for years (Corkin, Amaral, González, Johnson, & Hyman, 1997). As a result, his declarative memory was significantly affected, and he could not form new semantic knowledge. He lost the ability to form new memories, yet he could still remember information and events that had occurred prior to the surgery.

If the video above does not load, click here:  https://youtu.be/y0oyPi2pZAg

The Cerebellum and Prefrontal Cortex

Although the hippocampus seems to be more of a processing area for explicit memories, you could still lose it and be able to create implicit memories (procedural memory, motor learning, and classical conditioning), thanks to your  cerebellum  ( Figure M.20 ). For example, one classical conditioning experiment is to accustom subjects to blink when they are given a puff of air to the eyes. When researchers damaged the cerebellums of rabbits, they discovered that the rabbits were not able to learn the conditioned eye-blink response (Steinmetz, 1999; Green & Woodruff-Pak, 2000).

Other researchers have used brain scans, including positron emission tomography (PET) scans, to learn how people process and retain information. From these studies, it seems the prefrontal cortex is involved. In one study, participants had to complete two different tasks: either looking for the letter  a  in words (considered a perceptual task) or categorizing a noun as either living or non-living (considered a semantic task) (Kapur et al., 1994). Participants were then asked which words they had previously seen. Recall was much better for the semantic task than for the perceptual task. According to PET scans, there was much more activation in the left inferior prefrontal cortex in the semantic task. In another study, encoding was associated with left frontal activity, while retrieval of information was associated with the right frontal region (Craik et al., 1999).

If the video above does not load, click here:  https://youtu.be/_LIiYy-BRes

Neurotransmitters

There also appear to be specific neurotransmitters involved with the process of memory, such as epinephrine, dopamine, serotonin, glutamate, and acetylcholine (Myhrer, 2003). There continues to be discussion and debate among researchers as to which  neurotransmitter  plays which specific role (Blockland, 1996). Although we don’t yet know which role each neurotransmitter plays in memory, we do know that communication among neurones via neurotransmitters is critical for developing new memories. Repeated activity by neurones leads to increased neurotransmitters in the synapses and more efficient and more synaptic connections. This is how memory consolidation occurs.

It is also believed that strong emotions trigger the formation of strong memories, and weaker emotional experiences form weaker memories; this is called  arousal theory  (Christianson, 1992). For example, strong emotional experiences can trigger the release of neurotransmitters, as well as hormones, which strengthen memory; therefore, our memory for an emotional event is usually better than our memory for a non-emotional event. When humans and animals are stressed, the brain secretes more of the neurotransmitter glutamate, which helps them remember the stressful event (McGaugh, 2003). This is clearly evidenced by what is known as the flashbulb memory phenomenon.

A  flashbulb memory  is an exceptionally clear recollection of an important event ( Figure M.21 ). Where were you when you first heard about the 9/11 terrorist attacks? While you may be too young to remember this event, older people you know can most likely remember where they were and what they were doing. In fact, a Pew Research Center (2011) survey found that for those Americans who were age 8 or older at the time of the event, 97% can recall the moment they learned of this event, even a decade after it happened. It is important to understand that flashbulb memories are not only for negative events. These tend to be more universal examples, while extremely positive events are generally more personal making it difficult to give an example that everyone can relate to.

A photograph shows the World Trade Center buildings, shortly after two planes were flown into them on the morning of September 11, 2001. Thick, black clouds of smoke stream from both buildings.

Inaccurate and False Memories

Even flashbulb memories for important events can have decreased accuracy with the passage of time. For example, on at least three occasions, when asked how he heard about the terrorist attacks of 9/11, President George W. Bush responded inaccurately. In January 2002, less than 4 months after the attacks, the then sitting President Bush was asked how he heard about the attacks. He responded:

I was sitting there, and my Chief of Staff—well, first of all, when we walked into the classroom, I had seen this plane fly into the first building. There was a TV set on. And you know, I thought it was pilot error and I was amazed that anybody could make such a terrible mistake. (Greenberg, 2004, p. 2)

Contrary to what President Bush stated, no one saw the first plane hit, except people on the ground near the twin towers. Video footage of the first plane was not recorded because it was a normal Tuesday morning, until the first plane hit.

Memory is not like a video recording. Human memory, even flashbulb memories, can be frail. Different parts of them, such as the time, visual elements, and smells, are stored in different places. When something is remembered, these components have to be put back together for the complete memory, which is known as memory reconstruction. Each component creates a chance for an error to occur. False memory is remembering something that did not happen. Research participants have recalled hearing a word, even though they never heard the word (Roediger & McDermott, 2000).

Introduction to Psychology & Neuroscience Copyright © 2020 by Edited by Leanne Stevens is licensed under a Creative Commons Attribution 4.0 International License , except where otherwise noted.

Share This Book

  • Bipolar Disorder
  • Therapy Center
  • When To See a Therapist
  • Types of Therapy
  • Best Online Therapy
  • Best Couples Therapy
  • Best Family Therapy
  • Managing Stress
  • Sleep and Dreaming
  • Understanding Emotions
  • Self-Improvement
  • Healthy Relationships
  • Student Resources
  • Personality Types
  • Guided Meditations
  • Verywell Mind Insights
  • 2024 Verywell Mind 25
  • Mental Health in the Classroom
  • Editorial Process
  • Meet Our Review Board
  • Crisis Support

What Is Memory?

How memories help us

Kendra Cherry, MS, is a psychosocial rehabilitation specialist, psychology educator, and author of the "Everything Psychology Book."

presentation on human memory

Daniel B. Block, MD, is an award-winning, board-certified psychiatrist who operates a private practice in Pennsylvania.

presentation on human memory

Fabio / Getty Images

  • Organization

Why We Forget

How to improve memory.

  • How to Protect Memory

Memory refers to the psychological processes of acquiring, storing, retaining, and later retrieving information. There are three major processes involved in memory: encoding, storage, and retrieval.

Human memory involves the ability to both preserve and recover information. However, this is not a flawless process. Sometimes people forget or misremember things. Other times, information is not properly encoded in memory in the first place.

Memory problems are often relatively minor annoyances, like forgetting birthdays. However, they can also be a sign of serious conditions such as Alzheimer's disease  and other kinds of dementia . These conditions affect quality of life and ability to function.

This article discusses how memories are formed and why they are sometimes forgotten. It also covers the different types of memory and steps you can take to both improve and protect your memory.

How Memories Are Formed

In order to create a new memory, information must be changed into a usable form, which occurs through a process known as encoding . Once the information has been successfully encoded, it must be stored in memory for later use.

Researchers have long believed that memories form due to changes in brain neurons (nerve cells). Our understanding today is that memories are created through the connections that exist between these neurons—either by strengthening these connections or through the growth of new connections.

Changes in the connections between nerve cells (known as synapses ) are associated with the learning and retention of new information. Strengthening these connections helps commit information to memory.

This is why reviewing and rehearsing information improves the ability to remember it. Practice strengthens the connections between the synapses that store that memory.

Much of our stored memory lies outside of our awareness most of the time, except when we actually need to use it. The memory retrieval process allows us to bring stored memories into conscious awareness.

How Long Do Memories Last?

You can't discuss what memory is without also talking about how long memories last. Some memories are very brief, just seconds long, and allow people to take in sensory information about the world.

Short-term memories are a bit longer and last about 20 to 30 seconds. These memories mostly consist of the information people are currently focusing on and thinking about.

Some memories are capable of enduring much longer—lasting days, weeks, months, or even decades. Most of these long-term memories lie outside of immediate awareness but can be drawn into consciousness when needed.

Why Do We Remember Painful Memories?

Have you ever noticed that many times, painful memories tend to hang on for long periods of time? Research suggests that this is because of increased biological arousal during the negative experience, which increases the longevity of that memory.

Using Memory

To use the information that has been encoded into memory, it first has to be retrieved. There are many factors that can influence this process, including the type of information being used and the retrieval cues that are present.

Of course, this process is not always perfect. Have you ever felt like you had the answer to a question just out of your reach, for instance? This is an example of a perplexing memory retrieval issue known as lethologica or the tip-of-the-tongue phenomenon.

Organizing Memory

The ability to access and retrieve information from long-term memory allows us to actually use these memories to make decisions, interact with others, and solve problems . But in order to be retrievable, memories have to be organized in some way.

One way of thinking about memory organization is the semantic network model. This model suggests that certain triggers activate associated memories. Seeing or remembering a specific place might activate memories that have occurred in that location.

Thinking about a particular campus building, for example, might trigger memories of attending classes, studying, and socializing with peers.

Certain stimuli can also sometimes act as powerful triggers that draw memories into conscious awareness. Scent is one example. Smelling a particular smell, such as a perfume or fresh-baked cookies, can bring forth a rush of vivid memories connected to people and events from a person's past. 

In order to identify a scent, a person must remember when they have smelled it before, then connect it to visual information that occurred at the same time. So, when areas of the brain connected to memory are damaged, the ability to identify smells is actually impaired.

At the same time, researchers have found that scent can help trigger autobiographical memories in people who have Alzheimer's disease. This underscores just how powerful memories can be.

Types of Memory

While several different models of memory have been proposed, the stage model of memory is often used to explain the basic structure and function of memory. Initially proposed in 1968 by Richard Atkinson and Richard Shiffrin, this theory outlines three separate stages or types of memory : sensory memory, short-term memory, and long-term memory.

Sensory Memory

Sensory memory is the earliest stage of memory. During this stage, sensory information from the environment is stored for a very brief period of time, generally for no longer than a half-second for visual information and three or four seconds for auditory information.

People only pay attention to certain aspects of this sensory memory. Attending to sensory memory allows some of this information to pass into the next stage: short-term memory.

Short-Term Memory

Short-term memory, also known as active memory, is the information we are currently aware of or thinking about. In Freudian psychology, this memory would be referred to as the conscious mind . Paying attention to sensory memories generates information in short-term memory.

While many of our short-term memories are quickly forgotten, attending to this information allows it to continue to the next stage: long-term memory. Most of the information stored in active memory will be kept for approximately 20 to 30 seconds.

This capacity can be stretched somewhat by using memory strategies such as chunking , which involves grouping related information into smaller chunks.

The term "short-term memory" is often used interchangeably with "working memory," which refers to the processes that are used to temporarily store, organize, and manipulate information.

In a famous paper published in 1956, psychologist George Miller suggested that the capacity of short-term memory for storing a list of items was somewhere between five and nine. Some memory researchers now believe that the true capacity of short-term memory is probably closer to four.

Long-Term Memory

Long-term memory refers to the continuing storage of information. In Freudian psychology , long-term memory would be called the preconscious and unconscious .

This information is largely outside of our awareness but can be called into working memory to be used when needed. Some memories are fairly easy to recall, while others are much more difficult to access.

One model suggests that there are three main types of memory: sensory memory, short-term memory, and long-term memory. Sensory memory is very brief, short-term memory is slightly longer, and long-term memory can last a lifetime.

Forgetting is a surprisingly common event. Just consider how easy it is to forget someone’s name or overlook an important appointment. Why do people so often forget information they have learned in the past?

There are four basic explanations for why forgetting occurs :

  • Failure to store a memory
  • Interference
  • Motivated forgetting
  • Retrieval failure

Research has shown that one of the critical factors that influence memory failure is time. Information is often quickly forgotten, particularly if people do not actively review and rehearse the information.

Sometimes information is simply lost from memory and, in other cases, it was never stored correctly in the first place. Some memories compete with one another, making it difficult to remember certain information. In other instances, people actively try to forget things that they simply don’t want to remember.

No matter how great your memory is, there are probably a few things you can do to make it even better. Useful strategies to deal with mild memory loss include:

  • Write it down : The act of writing with a pen and paper helps implant the memory into your brain—and can also serve as a reminder or reference later on.
  • Attach meaning to it : You can remember something more easily if you attach meaning to it. For instance, if you associate a person you just meet with someone you already know, you may be able to remember their name better.
  • Repeat it : Repetition helps the memory become encoded beyond your short-term memory.
  • Group it : Information that is categorized becomes easier to remember and recall.
  • Test yourself : While it may seem like studying and rehearsing information is the best way to ensure that you will remember it, researchers have found that being tested on information is actually one of the best ways to improve recall .
  • Take a mental picture : Systematically trying to make a mental note of things you often forget (such as where you left your car keys) can help you remember things better.
  • Get enough rest : Research has also found that sleep plays a critical role in learning and the formation of new memories.
  • Use memorization techniques : Rehearsing information, employing mnemonics, and other memorization strategies can help combat minor memory problems.

Using strategies to boost memory can be helpful for recall and retention. By learning how to use these strategies effectively, you can sidestep the faulty areas of your memory and train your brain to function in new ways.

How to Protect Your Memory

While Alzheimer's disease and other age-related memory problems affect many older adults, the loss of memory during later adulthood might not inevitable. Certain abilities do tend to decline with age, but researchers have found that individuals in their 70s often perform just as well on many cognitive tests as those in their 20s.

By the time people reach their 80s, it is common to experience some decline in cognitive function. But some types of memory even increase with age.

To help protect your brain as you age, try some of these lifestyle strategies:

  • Avoid stress : Research has found that stress can have detrimental effects on areas of the brain associated with memory, including the hippocampus.
  • Avoid drugs, alcohol, and other neurotoxins : Drug use and excessive alcohol consumption have been linked to the deterioration of synapses (the connections between neurons). Exposure to dangerous chemicals such as heavy metals and pesticides can also have detrimental effects on the brain.
  • Get enough exercise : Regular physical activity helps improve oxygenation of the brain, which is vital for synaptic formation and growth.
  • Stimulate your brain : When it comes to memory, there is a lot of truth to the old adage of "use it or lose it." Researchers have found that people who have more mentally stimulating jobs are less likely to develop dementia.
  • Maintain a sense of self-efficacy : Having a strong sense of self-efficacy has been associated with maintaining good memory abilities during old age. Self-efficacy refers to the sense of control that people have over their own lives and destiny. A strong sense of self-efficacy has also been linked to lowered stress levels.

While there is no quick fix for ensuring that your memory stays intact as you age, researchers believe that avoiding stress, leading an active lifestyle, and remaining mentally engaged are important ways to decrease your risk of memory loss.

A Word From Verywell

Human memory is a complex process that researchers are still trying to better understand. Our memories make us who we are, yet the process is not perfect. While we are capable of remembering an astonishing amount of information, we are also susceptible to memory-related mistakes and errors.

National Institute of Neurological Disorders and Stroke. Brain basics: The life and death of a neuron .

Kark S, Kensinger E. Physiological arousal and visuocortical connectivity predict subsequent vividness of negative memories . Neurorep . 2019;30(12):800-804. doi:10.1097/WNR.0000000000001274

Zemla JC, Austerweil JL. Estimating semantic networks of groups and individuals from fluency data .  Comput Brain Behav . 2018;1(1):36-58. doi:10.1007/s42113-018-0003-7

Association for Psychological Science. Fragrant flashbacks .

Glachet O, El Haj M. Emotional and phenomenological properties of odor-evoked autobiographical memories in Alzheimer's disease .  Brain Sci . 2019;9(6):135. doi:10.3390/brainsci9060135

Camina E, Güell F. The neuroanatomical, neurophysiological and psychological basis of memory: Current models and their origins .  Front Pharmacol . 2017;8:438. doi:10.3389/fphar.2017.00438

Norris D. Short-term memory and long-term memory are still different . Psychol Bull . 2017;143(9):992-1009. doi:10.1037/bul0000108

Cowan N. The magical number 4 in short-term memory: a reconsideration of mental storage capacity . Behav Brain Sci . 2001;24(1):87-114. doi:10.1017/s0140525x01003922

Darby K, Sloutsky V. The cost of learning: Interference effects in memory development .  J Exp Psychol Gen . 2015;144(2):410-431. doi:10.1037/xge0000051

Mueller PA, Oppenheimer DM. The pen is mightier than the keyboard: Advantages of longhand over laptop note taking . Psychol Sci . 2014;25(6):1159-68. doi:10.1177/0956797614524581

McEwen BS, Nasca C, Gray JD. Stress effects on neuronal structure: Hippocampus, amygdala, and prefrontal cortex .  Neuropsychopharmacology . 2016;41(1):3-23. doi:10.1038/npp.2015.171

National Institutes of Health. A well-aged mind: Maintaining your cognitive health .

Goodman J, Packard MG. Memory systems and the addicted brain . Front Psychiatry . 2016;7. doi:10.3389/fpsyt.2016.00024

Mandolesi L, Polverino A, Montuori S, et al. Effects of physical exercise on cognitive functioning and wellbeing: Biological and psychological benefits .  Front Psychol . 2018;9:509. doi:10.3389/fpsyg.2018.00509

Kivimäki M, Walker KA, Pentti J, et al. Cognitive stimulation in the workplace, plasma proteins, and risk of dementia: three analyses of population cohort studies . BMJ . 2021;n1804. doi:10.1136/bmj.n1804

Beaudoin M, Desrichard O. Memory self-efficacy and memory performance in older adults: the mediating role of task persistence . Swiss J Psychol . 2017;76(1):23-33. doi:10.1024/1421-0185/a000188

Schönfeld P, Brailovskaia J, Zhang XC, Margraf J. Self-efficacy as a mechanism linking daily stress to mental health in students: a three-wave cross-lagged study . Psychol Rep . 2019;122(6):2074-2095. doi:10.1177/0033294118787496

By Kendra Cherry, MSEd Kendra Cherry, MS, is a psychosocial rehabilitation specialist, psychology educator, and author of the "Everything Psychology Book."

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • Front Hum Neurosci
  • PMC10410470

Cognitive neuroscience perspective on memory: overview and summary

Sruthi sridhar.

1 Department of Psychology, Mount Allison University, Sackville, NB, Canada

Abdulrahman Khamaj

2 Department of Industrial Engineering, College of Engineering, Jazan University, Jazan, Saudi Arabia

Manish Kumar Asthana

3 Department of Humanities and Social Sciences, Indian Institute of Technology Roorkee, Roorkee, India

4 Department of Design, Indian Institute of Technology Roorkee, Roorkee, India

Associated Data

The original contributions presented in this study are included in the article/supplementary material, further inquiries can be directed to the corresponding author.

This paper explores memory from a cognitive neuroscience perspective and examines associated neural mechanisms. It examines the different types of memory: working, declarative, and non-declarative, and the brain regions involved in each type. The paper highlights the role of different brain regions, such as the prefrontal cortex in working memory and the hippocampus in declarative memory. The paper also examines the mechanisms that underlie the formation and consolidation of memory, including the importance of sleep in the consolidation of memory and the role of the hippocampus in linking new memories to existing cognitive schemata. The paper highlights two types of memory consolidation processes: cellular consolidation and system consolidation. Cellular consolidation is the process of stabilizing information by strengthening synaptic connections. System consolidation models suggest that memories are initially stored in the hippocampus and are gradually consolidated into the neocortex over time. The consolidation process involves a hippocampal-neocortical binding process incorporating newly acquired information into existing cognitive schemata. The paper highlights the role of the medial temporal lobe and its involvement in autobiographical memory. Further, the paper discusses the relationship between episodic and semantic memory and the role of the hippocampus. Finally, the paper underscores the need for further research into the neurobiological mechanisms underlying non-declarative memory, particularly conditioning. Overall, the paper provides a comprehensive overview from a cognitive neuroscience perspective of the different processes involved in memory consolidation of different types of memory.

Introduction

Memory is an essential cognitive function that permits individuals to acquire, retain, and recover data that defines a person’s identity ( Zlotnik and Vansintjan, 2019 ). Memory is a multifaceted cognitive process that involves different stages: encoding, consolidation, recovery, and reconsolidation. Encoding involves acquiring and processing information that is transformed into a neuronal representation suitable for storage ( Liu et al., 2021 ; Panzeri et al., 2023 ). The information can be acquired through various channels, such as visual, auditory, olfactory, or tactile inputs. The acquired sensory stimuli are converted into a format the brain can process and retain. Different factors such as attention, emotional significance, and repetition can influence the encoding process and determine the strength and durability of the resulting memory ( Squire et al., 2004 ; Lee et al., 2016 ; Serences, 2016 ).

Consolidation includes the stabilization and integration of memory into long-term storage to increase resistance to interference and decay ( Goedert and Willingham, 2002 ). This process creates enduring structural modification in the brain and thereby has consequential effects on the function by reorganizing and strengthening neural connections. Diverse sources like sleep and stress and the release of neurotransmitters can influence memory consolidation. Many researchers have noted the importance of sleep due to its critical role in enabling a smooth transition of information from transient repositories into more stable engrams (memory traces) ( McGaugh, 2000 ; Clawson et al., 2021 ; Rakowska et al., 2022 ).

Retrieval involves accessing, selecting, and reactivating or reconstructing the stored memory to allow conscious access to previously encoded information ( Dudai, 2002 ). Retrieving memories depends on activating relevant neural pathways while reconstructing encoded information. Factors like contextual or retrieval cues and familiarity with the material can affect this process. Forgetting becomes a possibility if there are inadequate triggers for associated memory traces to activate upon recall. Luckily, mnemonic strategies and retrieval practice offer effective tools to enhance recovery rates and benefit overall memory performance ( Roediger and Butler, 2011 ).

Previous research implied that once a memory has been consolidated, it becomes permanent ( McGaugh, 2000 ; Robins, 2020 ). However, recent studies have found an additional phase called “reconsolidation,” during which stored memories, when reactivated, enter a fragile or liable state and become susceptible to modification or update ( Schiller et al., 2009 ; Asthana et al., 2015 ). The process highlights the notion that memory is not static but a dynamic system influenced by subsequent encounters. The concept of reconsolidation has much significance in memory modification therapies and interventions, as it offers a promising opportunity to target maladaptive or traumatic memories for modification specifically. However, more thorough investigations are needed to gain insight into the mechanisms and concrete implications of employing memory reconsolidation within therapeutic settings ( Bellfy and Kwapis, 2020 ).

The concept of memory is not reducible to a single unitary phenomenon; instead, evidence suggests that it can be subdivided into several distinct but interrelated constituent processes and systems ( Richter-Levin and Akirav, 2003 ). There are three major types of human memory: working memory, declarative memory (explicit), and non-declarative memory (implicit). All these types of memories involve different neural systems in the brain. Working memory is a unique transient active store capable of manipulating information essential for many complex cognitive operations, including language processing, reasoning, and judgment ( Atkinson and Shiffrin, 1968 ; Baddeley and Logie, 1999 ; Funahashi, 2017 ; Quentin et al., 2019 ). Previous models suggest the existence of three components that make up the working memory ( Baddeley and Hitch, 1974 ; Baddeley, 1986 ). One master component, the central executive, controls the two dependent components, the phonological loop (speech perception and language comprehension) and the visuospatial sketchpad (visual images and spatial impressions processing). Some models mention a third component known as the episodic buffer. It is theorized that the episodic buffer serves as an intermediary between perception, long-term memory, and two components of working memory (the phonological loop and visuospatial sketchpad) by storing integrated episodes or chunks from both sources ( Baddeley, 2000 ). Declarative memory (explicit memory) can be recalled consciously, including facts and events that took place in one’s life or information learned from books. It encompasses memories of both autobiographical experiences and memories associated with general knowledge. It is usually associated with the hippocampus–medial temporal lobe system ( Thompson and Kim, 1996 ; Ober, 2014 ). Non-declarative memory (implicit memory) refers to unconscious forms of learning such as skills, habits, and priming effects; this type of implicit learning does not involve conscious recollection but can include motor skill tasks that often require no thought prior to execution nor later recall upon completion. This type of memory usually involves the amygdala and other systems ( Thompson and Kim, 1996 ; Ober, 2014 ).

Working memory

Working memory is primarily associated with the prefrontal and posterior parietal cortex ( Sarnthein et al., 1998 ; Todd and Marois, 2005 ). Working memory is not localized to a single brain region, and research suggests that it is an emergent property arising from functional interactions between the prefrontal cortex (PFC) and the rest of the brain ( D’Esposito, 2007 ). Neuroimaging studies have explored the neural basis for the three components proposed by Baddeley and Hitch (1974) , the Central executive, the phonological loop, and the visuospatial sketch pad; there is evidence for the existence of a fourth component called the episodic buffer ( Baddeley, 2000 ).

The central executive plays a significant role in working memory by acting as the control center ( Shallice, 2002 ). It facilitates critical functions like attention allocation and coordination between the phonological loop and the visuospatial sketchpad ( Yu et al., 2023 ). Recent findings have illuminated the dual-functional network regulation, the cingulo-opercular network (CON) and the frontoparietal network (FPN), that underpins the central executive system ( Yu et al., 2023 ). The CON comprises the dorsal anterior cingulate cortex (dACC) and anterior insula (AI). In contrast, the FPN encompasses various regions, such as the dorsolateral prefrontal cortex (DLPFC) and frontal eye field (FEF), along with the intraparietal sulcus (IPS) ( Yu et al., 2023 ). Neuroimaging research has found evidence that elucidates the neural underpinnings of the executive attention control system to the dorsolateral prefrontal cortex (DLPFC) and the anterior cingulate cortex (ACC) ( Jung et al., 2022 ). The activation patterns indicate that the CON may have a broader top-down control function across the working memory process. At the same time, the FPN could be more heavily implicated in momentary control or processing at the trial level ( Yu et al., 2023 ). Evidence suggests that the central executive interacts with the phonological loop and visuospatial sketchpad to support working memory processes ( Baddeley, 2003 ; Buchsbaum, 2010 ; Menon and D’Esposito, 2021 ). The function, localization, and neural basis of this interaction are thought to involve the activation of specific brain regions associated with each component of working memory, as discussed in detail below.

The phonological loop is divided into two components: a storage system that maintains information (a few seconds) and a component involving subvocal rehearsal—which maintains and refreshes information in the working memory. Neuroanatomically, the phonological loop is represented in the Brodmann area (BA) 40 in the parietal cortex and the rehearsal components in BA 44 and 6, both situated in the frontal cortex ( Osaka et al., 2007 ). The left inferior frontal gyrus (Broca’s area) and the left posterior superior temporal gyrus (Wernicke’s area) has been proposed to play a critical role in supporting phonological and verbal working memory tasks, specifically the subvocal rehearsal system of the articulatory loop ( Paulesu et al., 1993 ; Buchsbaum et al., 2001 ; Perrachione et al., 2017 ). The phonological store in verbal short-term memory has been localized at the left supramarginal gyrus ( Graves et al., 2008 ; Perrachione et al., 2017 ).

Studies utilizing neuroimaging techniques have consistently yielded results indicating notable activation in these brain regions during phonological activities like recalling non-words and maintaining verbal information in memory ( Awh et al., 1996 ; Graves et al., 2008 ). During tasks that require phonological rehearsal, there was an increase in activation in the left inferior frontal gyrus ( Paulesu et al., 1993 ). Researchers have noted an increase in activity within the superior temporal gyrus-which plays a significant role in auditory processing-in individuals performing tasks that necessitate verbal information maintenance and manipulation ( Smith et al., 1998 ; Chein et al., 2003 ).

Additionally, lesion studies have provided further confirmation regarding the importance of these regions. These investigations have revealed that impairment in performing phonological working memory tasks can transpire following damage inflicted upon the left hemisphere, particularly on perisylvian language areas ( Koenigs et al., 2011 ). It is common for individuals with lesions affecting regions associated with the phonological loop, such as the left inferior frontal gyrus and superior temporal gyrus, to have difficulty performing verbal working memory tasks. Clinical cases involving patients diagnosed with aphasia and specific language impairments have highlighted challenges related to retaining and manipulating auditory information. For example, those who sustain damage specifically within their left inferior frontal gyrus often struggle with tasks involving phonological rehearsal and verbal working memory activities, and therefore, they tend to perform poorly in tasks that require manipulation or repetition of verbal stimuli ( Saffran, 1997 ; Caplan and Waters, 2005 ).

The visuospatial sketchpad is engaged in the temporary retention and manipulation of visuospatial facts, including mental pictures, spatial associations, and object placements ( Miyake et al., 2001 ). The visuospatial sketchpad is localized to the right hemisphere, including the occipital lobe, parietal and frontal areas ( Osaka et al., 2007 ). Ren et al. (2019) identified the localization of the visuospatial sketchpad, and these areas were the right infero-lateral prefrontal cortex, lateral pre-motor cortices, right inferior parietal cortex, and the dorsolateral occipital cortices ( Burbaud et al., 1999 ; Salvato et al., 2021 ). Moreover, the posterior parietal cortex and the intraparietal sulcus have been implicated in spatial working memory ( Xu and Chun, 2006 ). Additionally, some evidence is available for an increase in brain regions associated with the visuospatial sketchpad during tasks involving mental imagery and spatial processing. Neuroimaging studies have revealed increased neural activation in some regions of the parietal cortex, mainly the superior and posterior parietal cortex, while performing mental rotation tasks ( Cohen et al., 1996 ; Kosslyn et al., 1997 ). However, further research is needed to better understand the visuospatial working memory and its integration with other cognitive processes ( Baddeley, 2003 ). Lesions to the regions involving the visuospatial sketchpad can have detrimental effects on visuospatial working memory tasks. Individuals with lesions to the posterior parietal cortex may exhibit deficits in mental rotation tasks and may be unable to mentally manipulate the visuospatial representation ( Buiatti et al., 2011 ). Moreover, studies concerning lesions have shown that damage to the parietal cortex can result in short-term deficits in visuospatial memory ( Shafritz et al., 2002 ). Damage to the occipital cortex can lead to performance impairments in tasks that require the generation and manipulation of mental visual images ( Moro et al., 2008 ).

The fourth component of the working memory, termed episodic buffer, was proposed by Baddeley (2000) . The episodic buffer is a multidimensional but essentially passive store that can hold a limited number of chunks, store bound features, and make them available to conscious awareness ( Baddeley et al., 2010 ; Hitch et al., 2019 ). Although research has suggested that episodic buffer is localized to the hippocampus ( Berlingeri et al., 2008 ) or the inferior lateral parietal cortex, it is thought to be not dependent on a single anatomical structure but instead can be influenced by the subsystems of working memory, long term memory, and even through perception ( Vilberg and Rugg, 2008 ; Baddeley et al., 2010 ). The episodic buffer provides a crucial link between the attentional central executive and the multidimensional information necessary for the operation of working memory ( Baddeley et al., 2011 ; Gelastopoulos et al., 2019 ).

The interdependence of the working memory modules, namely the phonological loop and visuospatial sketchpad, co-relates with other cognitive processes, for instance, spatial cognition and attention allocation ( Repovs and Baddeley, 2006 ). It has been found that the prefrontal cortex (PFC) and posterior parietal cortex (PPC) have a crucial role in several aspects of spatial cognition, such as the maintenance of spatially oriented attention and motor intentions ( Jerde and Curtis, 2013 ). The study by Sellers et al. (2016) and the review by Ikkai and Curtis (2011) posits that other brain areas could use the activity in PFC and PPC as a guide and manifest outputs to guide attention allocation, spatial memory, and motor planning. Moreover, research indicates that verbal information elicits an activation response in the left ventrolateral prefrontal cortex (VLPFC) when retained in the phonological loop, while visuospatial information is represented by a corresponding level of activity within the right homolog region ( Narayanan et al., 2005 ; Wolf et al., 2006 ; Emch et al., 2019 ). Specifically, the study by Yang et al. (2022) investigated the roles of two regions in the brain, the right inferior frontal gyrus (rIFG) and the right supra-marginal gyrus (rSMG), as they relate to spatial congruency in visual working memory tasks. A change detection task with online repetitive transcranial magnetic stimulation applied concurrently at both locations during high visual WM load conditions determined that rIFG is involved in actively repositioning the location of objects. At the same time, rSMG is engaged in passive perception of the stability of the location of objects.

Recent academic studies have found evidence to support the development of a new working memory model known as the state-based model ( D’Esposito and Postle, 2015 ). This theoretical model proposes that the allocation of attention toward internal representations permits short-term retention within working memory ( Ghaleh et al., 2019 ). The state-based model consists of two main categories: activated LTM models and sensorimotor recruitment models; the former largely focuses upon symbolic stimuli categorized under semantic aspects, while the latter has typically been applied to more perceptual tasks in experiments. This framework posits that prioritization through regulating cognitive processes provides insight into various characteristics across different activity types, including capacity limitations, proactive interference, etcetera ( D’Esposito and Postle, 2015 ). For example, the paper by Ghaleh et al. (2019) provides evidence for two separate mechanisms involved in maintenance of auditory information in verbal working memory: an articulatory rehearsal mechanism that relies more heavily on left sensorimotor areas and a non-articulatory maintenance mechanism that critically relies on left superior temporal gyrus (STG). These findings support the state-based model’s proposal that attentional allocation is necessary for short-term retention in working memory.

State-based models were found to be consistent with the suggested storage mechanism as they do not require representation transfer from one dedicated buffer type; research has demonstrated that any population of neurons and synapses may serve as such buffers ( Maass and Markram, 2002 ; Postle, 2006 ; Avraham et al., 2017 ). The review by D’Esposito and Postle (2015) examined the evidence to determine whether a persistent neural activity, synaptic mechanisms, or a combination thereof support representations maintained during working memory. Numerous neural mechanisms have been hypothesized to support the short-term retention of information in working memory and likely operate in parallel ( Sreenivasan et al., 2014 ; Kamiński and Rutishauser, 2019 ).

Persistent neural activity is the neural mechanism by which information is temporarily maintained ( Ikkai and Curtis, 2011 ; Panzeri et al., 2023 ). Recent review by Curtis and Sprague (2021) has focused on the notion that persistent neural activity is a fundamental mechanism for memory storage and have provided two main arcs of explanation. The first arc, mainly underpinned by empirical evidence from prefrontal cortex (PFC) neurophysiology experiments and computational models, posits that PFC neurons exhibit sustained firing during working memory tasks, enabling them to store representations in their active state ( Thuault et al., 2013 ). Intrinsic persistent firing in layer V neurons in the medial PFC has been shown to be regulated by HCN1 channels, which contribute to the executive function of the PFC during working memory episodes ( Thuault et al., 2013 ). Additionally, research has also found that persistent neural firing could possibly interact with theta periodic activity to sustain each other in the medial temporal, prefrontal, and parietal regions ( Düzel et al., 2010 ; Boran et al., 2019 ). The second arc involves advanced neuroimaging approaches which have, more recently, enabled researchers to decode content stored within working memories across distributed regions of the brain, including parts of the early visual cortex–thus extending this framework beyond just isolated cortical areas such as the PFC. There is evidence that suggests simple, stable, persistent activity among neurons in stimulus-selective populations may be a crucial mechanism for sustaining WM representations ( Mackey et al., 2016 ; Kamiński et al., 2017 ; Curtis and Sprague, 2021 ).

Badre (2008) discussed the functional organization of the PFC. The paper hypothesized that the rostro-caudal gradient of a function in PFC supported a control hierarchy, whereas posterior to anterior PFC mediated progressively abstract, higher-order controls ( Badre, 2008 ). However, this outlook proposed by Badre (2008) became outdated; the paper by Badre and Nee (2018) presented an updated look at the literature on hierarchical control. This paper supports neither a unitary model of lateral frontal function nor a unidimensional abstraction gradient. Instead, separate frontal networks interact via local and global hierarchical structures to support diverse task demands. This updated perspective is supported by recent studies on the hierarchical organization of representations within the lateral prefrontal cortex (LPFC) and the progressively rostral areas of the LPFC that process/represent increasingly abstract information, facilitating efficient and flexible cognition ( Thomas Yeo et al., 2011 ; Nee and D’Esposito, 2016 ). This structure allows the brain to access increasingly abstract action representations as required ( Nee and D’Esposito, 2016 ). It is supported by fMRI studies showing an anterior-to-posterior activation movement when tasks become more complex. Anatomical connectivity between areas also supports this theory, such as Area 10, which has projections back down to Area 6 but not vice versa.

Finally, studies confirm that different regions serve different roles along a hierarchy leading toward goal-directed behavior ( Badre and Nee, 2018 ). The paper by Postle (2015) exhibits evidence of activity in the prefrontal cortex that reflects the maintenance of high-level representations, which act as top-down signals, and steer the circulation of neural pathways across brain networks. The PFC is a source of top-down signals that influence processing in the posterior and subcortical regions ( Braver et al., 2008 ; Friedman and Robbins, 2022 ). These signals either enhance task-relevant information or suppress irrelevant stimuli, allowing for efficient yet effective search ( D’Esposito, 2007 ; D’Esposito and Postle, 2015 ; Kerzel and Burra, 2020 ). The study by Ratcliffe et al. (2022) provides evidence of the dynamic interplay between executive control mechanisms in the frontal cortex and stimulus representations held in posterior regions for working memory tasks. Moreover, the review by Herry and Johansen (2014) discusses the neural mechanisms behind actively maintaining task-relevant information in order for a person to carry out tasks and goals effectively. This review of data and research suggests that working memory is a multi-component system allowing for both the storage and processing of temporarily active representations. Neural activity throughout the brain can be differentially enhanced or suppressed based on context through top-down signals emanating from integrative areas such as PFC, parietal cortex, or hippocampus to actively maintain task-relevant information when it is not present in the environment ( Herry and Johansen, 2014 ; Kerzel and Burra, 2020 ).

In addition, Yu et al. (2022) examined how brain regions from the ventral stream pathway to the prefrontal cortex were activated during working memory (WM) gate opening and closing. They defined gate opening as the switch from maintenance to updating and gate closing as the switch from updating to maintenance. The data suggested that cognitive branching increases during the WM gating process, thus correlating the gating process and an information approach to the PFC function. The temporal cortices, lingual gyrus (BA19), superior frontal gyri including frontopolar cortices, and middle and inferior parietal regions are involved in processes of estimating whether a response option available will be helpful for each case. During gate closing, on the other hand, medial and superior frontal regions, which have been associated with conflict monitoring, come into play, as well as orbitofrontal and dorsolateral prefrontal processing at later times when decreasing activity resembling stopping or downregulating cognitive branching has occurred, confirming earlier theories about these areas being essential for estimation of usefulness already stored within long-term memories ( Yu et al., 2022 ).

Declarative and non-declarative memory

The distinctions between declarative and non-declarative memory are often based on the anatomical features of medial temporal lobe regions, specifically those involving the hippocampus ( Squire and Zola, 1996 ; Squire and Wixted, 2011 ). In the investigation of systems implicated in the process of learning and memory formation, it has been posited that the participation of the hippocampus is essential for the acquisition of declarative memories ( Eichenbaum and Cohen, 2014 ). In contrast, a comparatively reduced level of hippocampal involvement may suffice for non-declarative memories ( Squire and Zola, 1996 ; Williams, 2020 ).

Declarative memory (explicit) pertains to knowledge about facts and events. This type of information can be consciously retrieved with effort or spontaneously recollected without conscious intention ( Dew and Cabeza, 2011 ). There are two types of declarative memory: Episodic and Semantic. Episodic memory is associated with the recollection of personal experiences. It involves detailed information about events that happened in one’s life. Semantic memory refers to knowledge stored in the brain as facts, concepts, ideas, and objects; this includes language-related information like meanings of words and mathematical symbol values along with general world knowledge (e.g., capitals of countries) ( Binder and Desai, 2011 ). The difference between episodic and semantic memory is that when one retrieves episodic memory, the experience is known as “remembering”; when one retrieves information from semantic memory, the experience is known as “knowing” ( Tulving, 1985 ; Dew and Cabeza, 2011 ). The hippocampus, medial temporal lobe, and the areas in the diencephalon are implicated in declarative memory ( Richter-Levin and Akirav, 2003 ; Derner et al., 2020 ). The ventral parietal cortex (VPC) is involved in declarative memory processes, specifically episodic memory retrieval ( Henson et al., 1999 ; Davis et al., 2018 ). The evidence suggests that VPC and hippocampus is involved in the retrieval of contextual details, such as the location and timing of the event, and the information is critical for the formation of episodic memory ( Daselaar, 2009 ; Hutchinson et al., 2009 ; Wiltgen et al., 2010 ). The prefrontal cortex (PFC) is involved in the encoding (medial PFC) and retrieval (lateral PFC) of declarative memories, specifically in the integration of information across different sensory modalities ( Blumenfeld and Ranganath, 2007 ; Li et al., 2010 ). Research also suggests that the amygdala may modulate other brain regions involved with memory processing, thus, contributing to an enhanced recall of negative or positive experiences ( Hamann, 2001 ; Ritchey et al., 2008 ; Sendi et al., 2020 ). Maintenance of the integrity of hippocampal circuitry is essential for ensuring that episodic memory, along with spatial and temporal context information, can be retained in short-term or long-term working memory beyond 15 min ( Ito et al., 2003 ; Rasch and Born, 2013 ). Moreover, studies have suggested that the amygdala plays a vital role in encoding and retrieving explicit memories, particularly those related to emotionally charged stimuli which are supported by evidence of correlations between hippocampal activity and amygdala modulation during memory formation ( Richter-Levin and Akirav, 2003 ; Qasim et al., 2023 ).

Current findings in neuroimaging studies assert that a vast array of interconnected brain regions support semantic memory ( Binder and Desai, 2011 ). This network merges information sourced from multiple senses alongside different cognitive faculties necessary for generating abstract supramodal views on various topics stored within our consciousness. Modality-specific sensory, motor, and emotional system within these brain regions serve specialized tasks like language comprehension, while larger areas of the brain, such as the inferior parietal lobe and most of the temporal lobe, participate in more generalized interpretation tasks ( Binder and Desai, 2011 ; Kuhnke et al., 2020 ). These regions lie at convergences of multiple perceptual processing streams, enabling increasingly abstract, supramodal representations of perceptual experience that support a variety of conceptual functions, including object recognition, social cognition, language, and the remarkable human capacity to remember the past and imagine the future ( Binder and Desai, 2011 ; Binney et al., 2016 ). The following section will discuss the processes underlying memory consolidation and storage within declarative memory.

Non-declarative (implicit) memories refer to unconscious learning through experience, such as habits and skills formed from practice rather than memorizing facts; these are typically acquired slowly and automatically in response to sensory input associated with reward structures or prior exposure within our daily lives ( Kesner, 2017 ). Non-declarative memory is a collection of different phenomena with different neural substrates rather than a single coherent system ( Camina and Güell, 2017 ). It operates by similar principles, depending on local changes to a circumscribed brain region, and the representation of these changes is unavailable to awareness ( Reber, 2008 ). Non-declarative memory encompasses a heterogenous collection of abilities, such as associative learning, skills, and habits (procedural memory), priming, and non-associative learning ( Squire and Zola, 1996 ; Camina and Güell, 2017 ). Studies have concluded that procedural memory for motor skills depends upon activity in diverse set areas such as the motor cortex, striatum, limbic system, and cerebellum; similarly, perceptual skill learning is thought to be associated with sensory cortical activation ( Karni et al., 1998 ; Mayes, 2002 ). Research suggests that mutual connections between brain regions that are active together recruit special cells called associative memory cells ( Wang et al., 2016 ; Wang and Cui, 2018 ). These cells help integrate, store, and remember related information. When activated, these cells trigger the recall of memories, leading to behaviors and emotional responses. This suggests that co-activated brain regions with these mutual connections are where associative memories are formed ( Wang et al., 2016 ; Wang and Cui, 2018 ). Additionally, observational data reveals that priming mechanisms within distinct networks, such as the “repetition suppression” effect observed in visual cortical areas associated with sensory processing and in the prefrontal cortex for semantic priming, are believed to be responsible for certain forms of conditioning and implicit knowledge transfer experiences exhibited by individuals throughout their daily lives ( Reber, 2008 ; Wig et al., 2009 ; Camina and Güell, 2017 ). However, further research is needed to better understand the mechanisms of consolidation in non-declarative memory ( Camina and Güell, 2017 ).

The process of transforming memory into stable, long-lasting from a temporary, labile memory is known as memory consolidation ( McGaugh, 2000 ). Memory formation is based on the change in synaptic connections of neurons representing the memory. Encoding causes synaptic Long-Term potentiation (LTP) or Long-Term depression (LTD) and induces two consolidation processes. The first is synaptic or cellular consolidation which involves remodeling synapses to produce enduring changes. Cellular consolidation is a short-term process that involves stabilizing the neural trace shortly after learning via structural brain changes in the hippocampus ( Lynch, 2004 ). The second is system consolidation, which builds on synaptic consolidation where reverberating activity leads to redistribution for long-term storage ( Mednick et al., 2011 ; Squire et al., 2015 ). System consolidation is a long-term process during which memories are gradually transferred to and integrated with cortical neurons, thus promoting their stability over time. In this way, memories are rendered less susceptible to forgetting. Hebb postulated that when two neurons are repeatedly activated simultaneously, they become more likely to exhibit a coordinated firing pattern of activity in the future ( Langille, 2019 ). This proposed enduring change in synchronized neuronal activation was consequently termed cellular consolidation ( Bermudez-Rattoni, 2010 ).

The following sections of this paper incorporate a more comprehensive investigation into various essential procedures connected with memory consolidation- namely: long-term potentiation (LTP), long-term depression (LTD), system consolidation, and cellular consolidation. Although these mechanisms have been presented briefly before this paragraph, the paper aims to offer greater insight into each process’s function within the individual capacity and their collective contribution toward memory consolidation.

Synaptic plasticity mechanisms implicated in memory stabilization

Long-Term Potentiation (LTP) and Long-Term Depression (LTP) are mechanisms that have been implicated in memory stabilization. LTP is an increase in synaptic strength, whereas LTD is a decrease in synaptic strength ( Ivanco, 2015 ; Abraham et al., 2019 ).

Long-Term Potentiation (LTP) is a phenomenon wherein synaptic strength increases persistently due to brief exposures to high-frequency stimulation ( Lynch, 2004 ). Studies of Long-Term Potentiation (LTP) have led to an understanding of the mechanisms behind synaptic strengthening phenomena and have provided a basis for explaining how and why strong connections between neurons form over time in response to stimuli.

The NMDA receptor-dependent LTP is the most commonly described LTP ( Bliss and Collingridge, 1993 ; Luscher and Malenka, 2012 ). In this type of LTP, when there is high-frequency stimulation, the presynaptic neuron releases glutamate, an excitatory neurotransmitter. Glutamate binds to the AMPA receptor on the postsynaptic neuron, which causes the neuron to fire while opening the NMDA receptor channel. The opening of an NMDA channel elicits a calcium ion influx into the postsynaptic neuron, thus initiating a series of phosphorylation events as part of the ensuing molecular cascade. Autonomously phosphorylated CaMKII and PKC, both actively functional through such a process, have been demonstrated to increase the conductance of pre-existing AMPA receptors in synaptic networks. Additionally, this has been shown to stimulate the introduction of additional AMPA receptors into synapses ( Malenka and Nicoll, 1999 ; Lynch, 2004 ; Luscher and Malenka, 2012 ; Bailey et al., 2015 ).

There are two phases of LTP: the early phase and the late phase. It has been established that the early phase LTP (E-LTP) does not require RNA or protein synthesis; therefore, its synaptic strength will dissipate in minutes if late LTP does not stabilize it. On the contrary, late-phase LTP (L-LTP) can sustain itself over a more extended period, from several hours to multiple days, with gene transcription and protein synthesis in the postsynaptic cell ( Frey and Morris, 1998 ; Orsini and Maren, 2012 ). The strength of presynaptic tetanic stimulation has been demonstrated to be a necessary condition for the activation of processes leading to late LTP ( Luscher and Malenka, 2012 ; Bailey et al., 2015 ). This finding is supported by research examining synaptic plasticity, notably Eric Kandel’s discovery that CREB–a transcription factor–among other cytoplasmic and nuclear molecules, are vital components in mediating molecular changes culminating in protein synthesis during this process ( Kaleem et al., 2011 ; Kandel et al., 2014 ). Further studies have shown how these shifts ultimately lead to AMPA receptor stabilization at post-synapses facilitating long-term potentiation within neurons ( Luscher and Malenka, 2012 ; Bailey et al., 2015 ).

The “synaptic tagging and capture hypothesis” explains how a weak event of tetanization at synapse A can transform to late-LTP if followed shortly by the strong tetanization of a different, nearby synapse on the same neuron ( Frey and Morris, 1998 ; Redondo and Morris, 2011 ; Okuda et al., 2020 ; Park et al., 2021 ). During this process, critical plasticity-related proteins (PRPs) are synthesized, which stabilize their own “tag” and that from the weaker synaptic activity ( Moncada et al., 2015 ). Recent evidence suggests that calcium-permeable AMPA receptors (CP-AMPARs) are involved in this form of heterosynaptic metaplasticity ( Park et al., 2018 ). The authors propose that the synaptic activation of CP-AMPARs triggers the synthesis of PRPs, which are then engaged by the weak induction protocol to facilitate LTP on the independent input. The paper also suggests that CP-AMPARs are required during the induction of LTP by the weak input for the full heterosynaptic metaplastic effect to be observed ( Park et al., 2021 ). Additionally, it has been further established that catecholamines such as dopamine plays an integral part in memory persistence by inducing PRP synthesis ( Redondo and Morris, 2011 ; Vishnoi et al., 2018 ). Studies have found that dopamine release in the hippocampus can enhance LTP and improve memory consolidation ( Lisman and Grace, 2005 ; Speranza et al., 2021 ).

Investigations into neuronal plasticity have indicated that synaptic strength alterations associated with certain forms of learning and memory may be analogous to those underlying Long-Term Potentiation (LTP). Research has corroborated this notion, demonstrating a correlation between these two phenomena ( Lynch, 2004 ). The three essential properties of Long-Term Potentiation (LTP) that have been identified are associativity, synapse specificity, and cooperativity ( Kandel and Mack, 2013 ). These characteristics provide empirical evidence for the potential role of LTP in memory formation processes. Specifically, associativity denotes the amplification of connections when weak stimulus input is paired with a powerful one; synapse specificity posits that this potentiating effect only manifests on synaptic locations exhibiting coincidental activity within postsynaptic neurons, while cooperativity suggests stimulated neuron needs to attain an adequate threshold of depolarization before LTP can be induced again ( Orsini and Maren, 2012 ).

There is support for the idea that memories are encoded by modification of synaptic strengths through cellular mechanisms such as LTP and LTD ( Nabavi et al., 2014 ). The paper by Nabavi et al. (2014) shows that fear conditioning, a type of associative memory, can be inactivated and reactivated by LTD and LTP, respectively. The findings of the paper support a causal link between these synaptic processes and memory. Moreover, the paper suggests that LTP is used to form neuronal assemblies that represent a memory, and LTD could be used to disassemble them and thereby inactivate a memory ( Nabavi et al., 2014 ). Hippocampal LTD has been found to play an essential function in regulating synaptic strength and forming memories, such as long-term spatial memory ( Ge et al., 2010 ). However, it is vital to bear in mind that studies carried out on LTP exceed those done on LTD; hence the literature on it needs to be more extensive ( Malenka and Bear, 2004 ; Nabavi et al., 2014 ).

Cellular consolidation and memory

For an event to be remembered, it must form physical connections between neurons in the brain, which creates a “memory trace.” This memory trace can then be stored as long-term memory ( Langille and Brown, 2018 ). The formation of a memory engram is an intricate process requiring neuronal depolarization and the influx of intracellular calcium ( Mank and Griesbeck, 2008 ; Josselyn et al., 2015 ; Xu et al., 2017 ). This initiation leads to a cascade involving protein transcription, structural and functional changes in neural networks, and stabilization during the quiescence period, followed by complete consolidation for its success. Interference from new learning events or disruption caused due to inhibition can abort this cycle leading to incomplete consolidation ( Josselyn et al., 2015 ).

Cyclic-AMP response element binding protein (CREB) has been identified as an essential transcription factor for memory formation ( Orsini and Maren, 2012 ). It regulates the expression of PRPs and enhances neuronal excitability and plasticity, resulting in changes to the structure of cells, including the growth of dendritic spines and new synaptic connections. Blockage or enhancement of CREB in certain areas can affect subsequent consolidation at a systems level–decreasing it prevents this from occurring, while aiding its presence allows even weak learning conditions to produce successful memory formation ( Orsini and Maren, 2012 ; Kandel et al., 2014 ).

Strengthening weakly encoded memories through the synaptic tagging and capture hypothesis may play an essential role in cellular consolidation. Retroactive memory enhancement has also been demonstrated in human studies, mainly when items are initially encoded with low strength but later paired with shock after consolidation ( Dunsmoor et al., 2015 ). The synaptic tagging and capture theory (STC) and its extension, the behavioral tagging hypothesis (BT), have both been used to explain synaptic specificity and the persistence of plasticity ( Moncada et al., 2015 ). STC proposed that electrophysiological activity can induce long-term changes in synapses, while BT postulates similar effects of behaviorally relevant neuronal events on learning and memory models. This hypothesis proposes that memory consolidation relies on combining two distinct processes: setting a “learning tag” and synthesizing plasticity-related proteins ( De novo protein synthesis, increased CREB levels, and substantial inputs to nearby synapses) at those tagged sites. BT explains how it is possible for event episodes with low-strength inputs or engagements can be converted into lasting memories ( Lynch, 2004 ; Moncada et al., 2015 ). Similarly, the emotional tagging hypothesis posits that the activation of the amygdala in emotionally arousing events helps to mark experiences as necessary, thus enhancing synaptic plasticity and facilitating transformation from transient into more permanent forms for encoding long-term memories ( Richter-Levin and Akirav, 2003 ; Zhu et al., 2022 ).

Cellular consolidation, the protein synthesis-dependent processes observed in rodents that may underlie memory formation and stabilization, has been challenging to characterize in humans due to the limited ability to study it directly ( Bermudez-Rattoni, 2010 ). Additionally, multi-trial learning protocols commonly used within human tests as opposed to single-trial experiments conducted with non-human subjects suggest there could be interference from subsequent information that impedes individual memories from being consolidated reliably. This raises important questions regarding how individuals can still form strong and long-lasting memories when exposed to frequent stimuli outside controlled laboratory conditions. Although this phenomenon remains undiscovered by science, it is of utmost significance for gaining a deeper understanding of our neural capacities ( Genzel and Wixted, 2017 ).

The establishment of distributed memory traces requires a narrow temporal window following the initial encoding process, during which cellular consolidation occurs ( Nader and Hardt, 2009 ). Once this period ends and consolidation has been completed, further protein synthesis inhibition or pharmacological disruption will be less effective at altering pre-existing memories and interfering with new learning due to the stabilization of the trace in its new neuronal network connections ( Nader and Hardt, 2009 ). Thus, systems consolidation appears critical for the long-term maintenance of memory within broader brain networks over extended periods after their formation ( Bermudez-Rattoni, 2010 ).

System consolidation and memory

Information is initially stored in both the hippocampus and neocortex ( Dudai et al., 2015 ). The hippocampus subsequently guides a gradual process of reorganization and stabilization whereby information present within the neocortex becomes autonomous from that in the hippocampal store. Scholars have termed this phenomenon “standard memory consolidation model” or “system consolidation” ( Squire et al., 2015 ).

The Standard Model suggests that information acquired during learning is simultaneously stored in both the hippocampus and multiple cortical modules. Subsequently, it posits that over a period of time which may range from weeks to months or longer, the hippocampal formation directs an integration process by which these various elements become enclosed into single unified structures within the cortex ( Gilboa and Moscovitch, 2021 ; Howard et al., 2022 ). These newly learned memories are then assimilated into existing networks without interference or compression when necessary ( Frankland and Bontempi, 2005 ). It is important to note that memory engrams already exist within cortical networks during encoding. They only need strengthening through links enabled by hippocampal assistance-overtime allowing remote memory storage without reliance on the latter structure. Data appears consistent across studies indicating that both AMPA-and NMDA receptor-dependent “tagging” processes occurring within the cortex are essential components of progressive rewiring, thus enabling longer-term retention ( Takeuchi et al., 2014 ; Takehara-Nishiuchi, 2020 ).

Recent studies have additionally demonstrated that the rate of system consolidation depends on an individual’s ability to relate new information to existing networks made up of connected neurons, popularly known as “schemas” ( Robin and Moscovitch, 2017 ). In situations where prior knowledge is present and cortical modules are already connected at the outset of learning, it has been observed that a hippocampal-neocortical binding process occurs similarly to when forming new memories ( Schlichting and Preston, 2015 ). The proposed framework involves the medial temporal lobe (MTL), which is involved in acquiring new information and binds different aspects of an experience into a single memory trace. In contrast, the medial prefrontal cortex (mPFC) integrates this information with the existing knowledge ( Zeithamova and Preston, 2010 ; van Kesteren et al., 2012 ). During consolidation and retrieval, MTL is involved in replaying memories to the neocortex, where they are gradually integrated with existing knowledge and schemas and help retrieve memory traces. During retrieval, the mPFC is thought to use existing knowledge and schemas to guide retrieval and interpretation of memory. This may involve the assimilation of newly acquired information into existing cognitive schemata as opposed to the comparatively slow progression of creating intercortical connections ( Zeithamova and Preston, 2010 ; van Kesteren et al., 2012 , 2016 ).

Medial temporal lobe structures are essential for acquiring new information and necessary for autobiographical (episodic) memory ( Brown et al., 2018 ). The consolidation of autobiographical memories depends on a distributed network of cortical regions. Brain areas such as entorhinal, perirhinal, and parahippocampal cortices are essential for learning new information; however, they have little impact on the recollection of the past ( Squire et al., 2015 ). The hippocampus is a region of the brain that forms episodic memories by linking multiple events to create meaningful experiences ( Cooper and Ritchey, 2019 ). It receives information from all areas of the association cortex and cingulate cortex, subcortical regions via the fornix, as well as signals originating within its entorhinal cortex (EC) and amygdala regarding emotionally laden or potentially hazardous stimuli ( Sorensen, 2009 ). Such widespread connectivity facilitates the construction of an accurate narrative underpinning each remembered episode, transforming short-term into long-term recollections ( Richter-Levin and Akirav, 2003 ).

Researchers have yet to establish a consensus regarding where semantic memory information is localized within the brain ( Roldan-Valadez et al., 2012 ). Some proponents contend that such knowledge is lodged within perceptual and motor systems, triggered when we initially associate with a given object. This point of view is supported by studies highlighting how neural activity occurs initially in the occipital cortex, followed by left temporal lobe involvement during processing and pertinent contributions to word selection/retrieval via activation of left inferior frontal cortices ( Patterson et al., 2007 ). Moreover, research indicates elevated levels of fusiform gyrus engagement (a ventral surface region encompassing both temporal lobes) occurring concomitantly with verbal comprehension initiatives, including reading and naming tasks ( Patterson et al., 2007 ).

Research suggests that the hippocampus is needed for a few years after learning to support semantic memory (factual information), yet, it is not needed for the long term ( Squire et al., 2015 ). However, some forms of memory remain dependent on the hippocampus, such as the retrieval of spatial memory ( Wiltgen et al., 2010 ). Similarly, the Multiple-trace theory ( Moscovitch et al., 2006 ), also known as the transformation hypothesis ( Winocur and Moscovitch, 2011 ), posits that hippocampal engagement is necessary for memories that retain contextual detail such as episodic memories. Consolidation of memories into the neocortex is theorized to involve a loss of specific finer details, such as temporal and spatial information, in addition to contextual elements. This transition ultimately results in an evolution from episodic memory toward semantic memory, which consists mainly of gist-based facts ( Moscovitch et al., 2006 ).

Sleep and memory consolidation

Sleep is an essential physiological process crucial to memory consolidation ( Siegel, 2001 ). Sleep is divided into two stages: Non-rapid Eye Movement (NREM) sleep and Rapid Eye Movement (REM) sleep. NREM sleep is divided into three stages: N1, N2, and N3 (AKA Slow Wave Sleep or SWS) ( Rasch and Born, 2013 ). Each stage displays unique oscillatory patterns and phenomena responsible for consolidating memories in distinct ways. The first stage, or N1 sleep, is when an individual transitions between wakefulness and sleep. This type of sleep is characterized by low-amplitude, mixed-frequency brain activity. N1 sleep is responsible for the initial encoding of memories ( Rasch and Born, 2013 ). The second stage, or N2 sleep, is characterized by the occurrence of distinct sleep spindles and K-complexes in EEG. N2 is responsible for the consolidation of declarative memories ( Marshall and Born, 2007 ). The third stage of sleep N3, also known as slow wave sleep (SWS), is characterized by low-frequency brain activity, slow oscillations, and high amplitude. The slow oscillations which define the deepest stage of sleep are trademark rhythms of NREM sleep. These slow oscillations are delta waves combined to indicate slow wave activity (SWA), which is implicated in memory consolidation ( Tononi and Cirelli, 2003 ; Stickgold, 2005 ; Kim et al., 2019 ). Sleep spindles are another trademark defining NREM sleep ( Stickgold, 2005 ). Ripples are high-frequency bursts, and when combined with irregularly occurring sharp waves (high amplitude), they form the sharp-wave ripple (SWR). These spindles and the SWRs coordinate the reactivation and redistribution of hippocampus-dependent memories to neocortical sites ( Ngo et al., 2020 ; Girardeau and Lopes-dos-Santos, 2021 ). The third stage is also responsible for the consolidation of procedural memories, such as habits and motor skills ( Diekelmann and Born, 2010 ). During SWS, there is minimal cholinergic activity and intermediate noradrenergic activity ( Datta and MacLean, 2007 ).

Finally, the fourth stage of sleep is REM sleep, characterized by phasic REMs and muscle atonia ( Reyes-Resina et al., 2021 ). During REM sleep, there is high cholinergic activity, serotonergic and noradrenergic activity are at a minimum, and high theta activity ( Datta and MacLean, 2007 ). REM sleep is also characterized by local increases in plasticity-related immediate-early gene activity, which might favor the subsequent synaptic consolidation of memories in the cortex ( Ribeiro, 2007 ; Diekelmann and Born, 2010 ; Reyes-Resina et al., 2021 ). The fourth stage of sleep is responsible for the consolidation of emotional memories and the integration of newly acquired memories into existing knowledge structures ( Rasch and Born, 2013 ). Studies indicate that the cholinergic system plays an imperative role in modifying these processes by toggling the entire thalamo-cortico-hippocampal network between distinct modes, namely high Ach encoding mode during active wakefulness and REM sleep and low Ach consolidation mode during quiet wakefulness and NREM sleep ( Bergmann and Staresina, 2017 ; Li et al., 2020 ). Consequently, improving neocortical hippocampal communication results in efficient memory encoding/synaptic plasticity, whereas hippocampo-neocortical interactions favor better systemic memory consolidation ( Diekelmann and Born, 2010 ).

The dual process hypothesis of memory consolidation posits that SWS facilitates declarative, hippocampus-dependent memory, whereas REM sleep facilitates non-declarative hippocampus-independent memory ( Maquet, 2001 ; Diekelmann and Born, 2010 ). On the other hand, the sequential hypothesis states that different sleep stages play a sequential role in memory consolidation. Memories are encoded during wakefulness, consolidated during NREM sleep, and further processed and integrated during REM sleep ( Rasch and Born, 2013 ). However, there is evidence present that contradicts the sequential hypothesis. A study by Goerke et al. (2013) found that declarative memories can be consolidated during REM sleep, suggesting that the relationship between sleep stages and memory consolidation is much more complex than a sequential model. Moreover, other studies indicate the importance of coordinating specific sleep phases with learning moments for optimal memory retention. This indicates that the timing of sleep has more influence than the specific sleep stages ( Gais et al., 2006 ). The active system consolidation theory suggests that an active consolidation process results from the selective reactivation of memories during sleep; the brain selectively reactivates newly encoded memories during sleep, which enhances and integrates them into the network of pre-existing long-term memories ( Born et al., 2006 ; Howard et al., 2022 ). Research has suggested that slow-wave sleep (SWS) and rapid eye movement (REM) sleep have complementary roles in memory consolidation. Declarative and non-declarative memories benefiting differently depending on which sleep stage they rely on ( Bergmann and Staresina, 2017 ). Specifically, during SWS, the brain actively reactivates and reorganizes hippocampo-neocortical memory traces as part of system consolidation. Following this, REM sleep is crucial for stabilizing these reactivated memory traces through synaptic consolidation. While SWS may initiate early plastic processes in hippocampo-neocortical memory traces by “tagging” relevant neocortico-neocortical synapses for later consolidation ( Frey and Morris, 1998 ), long-term plasticity requires subsequent REM sleep ( Rasch and Born, 2007 , 2013 ).

The active system consolidation hypothesis is not the only mechanism proposed for memory consolidation during sleep. The synaptic homeostasis hypothesis proposes that sleep is necessary for restoring synaptic homeostasis, which is challenged by synaptic strengthening triggered by learning during wake and synaptogenesis during development ( Tononi and Cirelli, 2014 ). The synaptic homeostasis hypothesis assumes consolidation is a by-product of the global synaptic downscaling during sleep ( Puentes-Mestril and Aton, 2017 ). The two models are not mutually exclusive, and the hypothesized processes probably act in concert to optimize the memory function of sleep ( Diekelmann and Born, 2010 ).

Non-rapid eye movement sleep plays an essential role in the systems consolidation of memories, with evidence showing that different oscillations are involved in this process ( Düzel et al., 2010 ). With an oscillatory sequence initiated by a slow frontal cortex oscillation (0.5–1 Hz) traveling to the medial temporal lobe and followed by a sharp-wave ripple (SWR) in the hippocampus (100–200 Hz). Replay activity of memories can be measured during this oscillatory sequence across various regions, including the motor cortex and visual cortex ( Ji and Wilson, 2006 ; Eichenlaub et al., 2020 ). Replay activity of memory refers to the phenomenon where the hippocampus replays previously experienced events during sharp wave ripples (SWRs) and theta oscillations ( Zielinski et al., 2018 ). During SWRs, short, transient bursts of high-frequency oscillations occur in the hippocampus. During theta oscillations, hippocampal spikes are ordered according to the locations of their place fields during behavior. These sequential activities are thought to play a role in memory consolidation and retrieval ( Zielinski et al., 2018 ). The paper by Zielinski et al. (2018) suggests that coordinated hippocampal-prefrontal representations during replay and theta sequences play complementary and overlapping roles at different stages in learning, supporting memory encoding and retrieval, deliberative decision-making, planning, and guiding future actions.

Additionally, the high-frequency oscillations of SWR reactivate groups of neurons attributed to spatial information encoding to align synchronized activity across an array of neural structures, which results in distributed memory creation ( Swanson et al., 2020 ; Girardeau and Lopes-dos-Santos, 2021 ). Parallel to this process is slow oscillation or slow-wave activity within cortical regions, which reflects synced neural firing and allows regulation of synaptic weights, which is in accordance with the synaptic homeostasis hypothesis (SHY). The SHY posits that downscaling synaptic strengths help incorporate new memories by avoiding saturation of resources during extended periods–features validated by discoveries where prolonged wakefulness boosts amplitude while it diminishes during stretches of enhanced sleep ( Girardeau and Lopes-dos-Santos, 2021 ).

During REM sleep, the brain experiences “paradoxical” sleep due to the similarity in activity to wakefulness. This stage plays a significant role in memory processing. Theta oscillations which are dominant during REM sleep, are primarily observed in the hippocampus, and these are involved in memory consolidation ( Landmann et al., 2014 ). There has been evidence of coherence between theta oscillations in the hippocampus, medial frontal cortex, and amygdala, which support their involvement in memory consolidation ( Popa et al., 2010 ). During REM sleep, phasic events such as ponto-geniculo-occipital waves originating from the brainstem coordinate activity across various brain structures and may contribute to memory consolidation processes ( Rasch and Born, 2013 ). Research has suggested that sleep-associated consolidation may be mediated by the degree of overlap between new and already known material whereby, if the acquired information is similar to the information one has learned, it is more easily consolidated during sleep ( Tamminen et al., 2010 ; Sobczak, 2017 ).

In conclusion, understanding more about how the brains cycle through different stages of sleep, including specific wave patterns, offers valuable insight into the ability to store memories effectively. While NREM sleep is associated with SWRs and slow oscillations, facilitating memory consolidation and synaptic downscaling, REM sleep, characterized by theta oscillations and phasic events, contributes to memory reconsolidation and the coordination of activity across brain regions. By exploring the interactions between sleep stages, oscillations, and memory processes, one may learn more about how sleep impacts brain function and cognition in greater detail.

Century has passed since we addressed memory, and several notable findings have moved from bench-to-bedside research. Several cross-talks between multidiscipline have been encouraged. Nevertheless, further research is needed into neurobiological mechanisms of non-declarative memory, such as conditioning ( Gallistel and Balsam, 2014 ). Modern research indicates that structural change that encodes information is likely at the level of the synapse, and the computational mechanisms are implemented at the level of neural circuitry. However, it also suggests that intracellular mechanisms realized at the molecular level, such as micro RNAs, should not be discounted as potential mechanisms. However, further research is needed to study the molecular and structural changes brought on by implicit memory ( Gallistel and Balsam, 2014 ).

The contribution of non-human animal studies toward our understanding of memory processes cannot be understated; hence recognizing their value is vital for moving forward. While this paper predominantly focused on cognitive neuroscience perspectives, some articles cited within this paper were sourced from non-human animal studies providing fundamental groundwork and identification of critical mechanisms relevant to human memories. A need persists for further investigation—primarily with humans—which can validate existing findings from non-human animals. Moving forward, it is prudent for researchers to bridge the gap between animal and human investigations done while exploring parallels and exploring unique aspects of human memory processes. By integrating findings from both domains, one can gain a more comprehensive understanding of the complexities of memory and its underlying neural mechanisms. Such investigations will broaden the horizon of our memory process and answer the complex nature of memory storage.

This paper attempted to provide an overview and summarize memory and its processes. The paper focused on bringing the cognitive neuroscience perspective on memory and its processes. This may provide the readers with the understanding, limitations, and research perspectives of memory mechanisms.

Data availability statement

Author contributions.

SS and MKA: conceptualization, framework, and manuscript writing. AK: review and editing of the manuscript. All authors contributed to the article and approved the submitted version.

Acknowledgments

We gratefully thank students and Indian Institute of Technology Roorkee (IITR) office staff for their conditional and unconditional support. We also thank the Memory and Anxiety Research Group (MARG), IIT Roorkee for its constant support.

Funding Statement

MKA was supported by the F.I.G. grant (IITR/SRIC/2741). The funding agency had no role in the preparation of the manuscript.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

  • Abraham W. C., Jones O. D., Glanzman D. L. (2019). Is plasticity of synapses the mechanism of long-term memory storage? NPJ Sci. Learn. 4 : 9 . 10.1038/s41539-019-0048-y [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Asthana M. K., Brunhuber B., Mühlberger A., Reif A., Schneider S., Herrmann M. J. (2015). Preventing the return of fear using reconsolidation update mechanisms depends on the met-allele of the brain derived neurotrophic factor val66met polymorphism. Int. J. Neuropsychopharmacol . 19 : yv137 . 10.1093/ijnp/pyv137 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Atkinson R. C., Shiffrin R. M. (1968). “ Human memory: a proposed system and its control processes ,” in Psychology of learning and motivation , Vol. 2 eds Spence K. W., Spence J. T. (Amsterdam: Elsevier; ), 89–195. 10.1016/S0079-7421(08)60422-3 [ CrossRef ] [ Google Scholar ]
  • Avraham G., Leib R., Pressman A., Simo L. S., Karniel A., Shmuelof L., et al. (2017). State-based delay representation and its transfer from a game of pong to reaching and tracking. Eneuro 4 ENEURO.179–ENEURO.117. 10.1523/eneuro.0179-17.2017 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Awh E., Jonides J., Smith E. E., Schumacher E. H., Koeppe R. A., Katz S. (1996). Dissociation of storage and rehearsal in verbal working memory: evidence From Positron Emission Tomography. Psychol. Sci. 7 25–31. 10.1111/j.1467-9280.1996.tb00662.x [ CrossRef ] [ Google Scholar ]
  • Baddeley A. (2000). The episodic buffer: a new component of working memory? Trends Cogn. Sci. 4 417–423. 10.1016/S1364-6613(00)01538-2 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Baddeley A. (2003). Working memory: looking back and looking forward. Nat. Rev. Neurosci. 4 829–839. 10.1038/nrn1201 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Baddeley A. D. (1986). Working memory. Oxford: Oxford University Press. [ Google Scholar ]
  • Baddeley A. D., Allen R. J., Hitch G. J. (2011). Binding in visual working memory: the role of the episodic buffer. Neuropsychologia 49 1393–1400. 10.1016/j.neuropsychologia.2010.12.042 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Baddeley A. D., Hitch G. (1974). “ Working memory ,” in Psychology of learning and motivation , Vol. 8 ed. Bower G. A. (Amsterdam: Elsevier; ), 47–89. 10.1016/S0079-7421(08)60452-1 [ CrossRef ] [ Google Scholar ]
  • Baddeley A. D., Logie R. H. (1999). “ Working memory: the multiple-component model ,” in Models of working memory , 1st Edn, eds Miyake A., Shah P. (Cambridge: Cambridge University Press; ), 28–61. 10.1017/CBO9781139174909.005 [ CrossRef ] [ Google Scholar ]
  • Baddeley A., Allen R. J., Hitch G. J. (2010). Investigating the episodic buffer. Psychol. Belgica 50 223 . 10.5334/pb-50-3-4-223 [ CrossRef ] [ Google Scholar ]
  • Badre D. (2008). Cognitive control, hierarchy, and the Rostro–caudal organization of the frontal lobes. Trends Cogn. Sci. 12 193–200. 10.1016/j.tics.2008.02.004 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Badre D., Nee D. E. (2018). Frontal cortex and the hierarchical control of behavior. Trends Cogn. Sci. 22 170–188. 10.1016/j.tics.2017.11.005 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bailey C. H., Kandel E. R., Harris K. M. (2015). Structural components of synaptic plasticity and memory consolidation. Cold Spring Harb. Perspect. Biol. 7 : a021758 . 10.1101/cshperspect.a021758 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bellfy L., Kwapis J. L. (2020). Molecular mechanisms of reconsolidation-dependent memory updating. Int. J. Mol. Sci. 21 : 6580 . 10.3390/ijms21186580 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bergmann T. O., Staresina B. P. (2017). “ Neuronal oscillations and reactivation subserving memory consolidation ,” in Cognitive neuroscience of memory consolidation , eds Axmacher N., Rasch B. (Cham: Springer; ), 185–207. 10.1007/978-3-319-45066-7_12 [ CrossRef ] [ Google Scholar ]
  • Berlingeri M., Bottini G., Basilico S., Silani G., Zanardi G., Sberna M., et al. (2008). Anatomy of the episodic buffer: a Voxel-based morphometry study in patients with dementia. Behav. Neurol. 19 29–34. 10.1155/2008/828937 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bermudez-Rattoni F. (2010). Is memory consolidation a multiple-circuit system? Proc. Natl. Acad. Sci. U.S.A. 107 8051–8052. 10.1073/pnas.1003434107 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Binder J. R., Desai R. H. (2011). The neurobiology of semantic memory. Trends Cogn. Sci. 15 527–536. 10.1016/j.tics.2011.10.001 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Binney R. J., Hoffman P., Lambon Ralph M. A. (2016). Mapping the multiple graded contributions of the anterior temporal lobe representational hub to abstract and social concepts: evidence from distortion-corrected fmri. Cereb. Cortex 26 4227–4241. 10.1093/cercor/bhw260 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Bliss T. V., Collingridge G. L. (1993). A synaptic model of memory: long-term potentiation in the hippocampus. Nature 361 31–39. 10.1038/361031a0 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Blumenfeld R. S., Ranganath C. (2007). Prefrontal cortex and long-term memory encoding: an integrative review of findings from Neuropsychology and Neuroimaging. Neuroscientist 13 280–291. 10.1177/1073858407299290 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Boran E., Fedele T., Klaver P., Hilfiker P., Stieglitz L., Grunwald T., et al. (2019). Persistent hippocampal neural firing and hippocampal-cortical coupling predict verbal working memory load. Sci. Adv. 5 : eaav3687 . 10.1126/sciadv.aav3687 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Born J., Rasch B., Gais S. (2006). Sleep to remember. Neuroscientist 12 410–424. 10.1177/1073858406292647 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Braver T. S., Gray J. R., Burgess G. C. (2008). “ Explaining the many varieties of working memory variation: dual mechanisms of cognitive control ,” in Variation in working memory , 1st Edn, eds Conway A., Jarrold C., Kane M., Miyake A., Towse J. (New York, NY: Oxford University Press; ), 76–106. 10.1093/acprof:oso/9780195168648.003.0004 [ CrossRef ] [ Google Scholar ]
  • Brown T. I., Rissman J., Chow T. E., Uncapher M. R., Wagner A. D. (2018). Differential medial temporal lobe and parietal cortical contributions to real-world autobiographical episodic and autobiographical semantic memory. Sci. Rep. 8 : 6190 . 10.1038/s41598-018-24549-y [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Buchsbaum B. R. (2010). Neural basis of working memory. Encyclopedia of Behavioral Neuroscience 334–341. 10.1016/b978-0-08-045396-5.00161-5 [ CrossRef ] [ Google Scholar ]
  • Buchsbaum B. R., Hickok G., Humphries C. (2001). Role of left posterior superior temporal gyrus in phonological processing for speech perception and production. Cogn. Sci. 25 663–678. 10.1207/s15516709cog2505_2 [ CrossRef ] [ Google Scholar ]
  • Buiatti T., Mussoni A., Toraldo A., Skrap M., Shallice T. (2011). Two qualitatively different impairments in making rotation operations. Cortex 47 166–179. 10.1016/j.cortex.2009.10.006 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Burbaud P., Camus O., Caillé J., Biolac B., Allard M. (1999). Influence of individual strategies on brain activation patterns during cognitive tasks. J. Neuroradiol. 26 59–65. [ PubMed ] [ Google Scholar ]
  • Camina E., Güell F. (2017). The neuroanatomical, neurophysiological and psychological basis of memory: current models and their origins. Front. Pharmacol. 8 : 438 . 10.3389/fphar.2017.00438 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Caplan D., Waters G. (2005). The relationship between age, processing speed, working memory capacity, and language comprehension. Memory 13 403–413. 10.1080/09658210344000459 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Chein J. M., Ravizza S. M., Fiez J. A. (2003). Using neuroimaging to evaluate models of working memory and their implications for language processing. J. Neurolinguist. 16 315–339. 10.1016/s0911-6044(03)00021-6 [ CrossRef ] [ Google Scholar ]
  • Clawson B. C., Pickup E. J., Ensing A., Geneseo L., Shaver J., Gonzalez-Amoretti J., et al. (2021). Causal role for sleep-dependent reactivation of learning-activated sensory ensembles for fear memory consolidation . Nat. Commun . 12 , 1–13. 10.1038/s41467-021-21471-2 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cohen M. S., Kosslyn S. M., Breiter H. C., DiGirolamo G. J., Thompson W. L., Anderson A. K., et al. (1996). Changes in cortical activity during mental rotation a mapping study using functional MRI. Brain 119 89–100. 10.1093/brain/119.1.89 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cooper R. A., Ritchey M. (2019). Progression from feature-specific brain activity to hippocampal binding during episodic encoding. J. Neurosci. 40 1701–1709. 10.1523/jneurosci.1971-19.2019 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Curtis C. E., Sprague T. C. (2021). Persistent activity during working memory from front to back. Front. Neural Circ. 15 : 696060 . 10.3389/fncir.2021.696060 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • D’Esposito M. (2007). From cognitive to neural models of working memory. Philos. Trans. R. Soc. B Biol. Sci. 362 761–772. 10.1098/rstb.2007.2086 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • D’Esposito M., Postle B. R. (2015). The cognitive neuroscience of working memory. Annu. Rev. Psychol. 66 115–142. 10.1146/annurev-psych-010814-015031 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Daselaar S. M. (2009). Posterior midline and ventral parietal activity is associated with retrieval success and encoding failure. Front. Hum. Neurosci. 3 : 13 . 10.3389/neuro.09.013.2009 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Datta S., MacLean R. R. (2007). Neurobiological mechanisms for the regulation of mammalian sleep–wake behavior: reinterpretation of historical evidence and inclusion of contemporary cellular and molecular evidence. Neurosci. Biobehav. Rev. 31 775–824. 10.1016/j.neubiorev.2007.02.004 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Davis S. W., Wing E. A., Cabeza R. (2018). Contributions of the ventral parietal cortex to declarative memory. Handb. Clin. Neurol. 151 525–553. 10.1016/b978-0-444-63622-5.00027-9 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Derner M., Dehnen G., Chaieb L., Reber T. P., Borger V., Surges R., et al. (2020). Patterns of single-neuron activity during associative recognition memory in the human medial temporal lobe. Neuroimage 221 : 117214 . 10.1016/j.neuroimage.2020.117214 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Dew I. T. Z., Cabeza R. (2011). The porous boundaries between explicit and implicit memory: behavioral and neural evidence. Ann. N. Y. Acad. Sci. 1224 174–190. 10.1111/j.1749-6632.2010.05946.x [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Diekelmann S., Born J. (2010). The memory function of sleep. Nat. Rev. Neurosci. 11 114–126. 10.1038/nrn2762 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Dudai Y. (2002). Molecular bases of long-term memories: a question of persistence. Curr. Opin. Neurobiol. 12 211–216. 10.1016/s0959-4388(02)00305-7 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Dudai Y., Karni A., Born J. (2015). The consolidation and transformation of memory. Neuron 88 20–32. 10.1016/j.neuron.2015.09.004 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Dunsmoor J. E., Murty V. P., Davachi L., Phelps E. A. (2015). Emotional learning selectively and retroactively strengthens memories for related events. Nature 520 345–348. 10.1038/nature14106 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Düzel E., Penny W. D., Burgess N. (2010). Brain oscillations and memory. Curr. Opin. Neurobiol. 20 143–149. 10.1016/j.conb.2010.01.004 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Eichenbaum H., Cohen N. J. (2014). Can we reconcile the declarative memory and spatial navigation views on hippocampal function? Neuron 83 764–770. 10.1016/j.neuron.2014.07.032 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Eichenlaub J.-B., Jarosiewicz B., Saab J., Franco B., Kelemen J., Halgren E., et al. (2020). Replay of learned neural firing sequences during rest in human motor cortex. Cell Rep. 31 : 107581 . 10.1016/j.celrep.2020.107581 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Emch M., von Bastian C. C., Koch K. (2019). Neural correlates of verbal working memory: an fMRI meta-analysis. Front. Hum. Neurosci. 13 : 180 . 10.3389/fnhum.2019.00180 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Frankland P. W., Bontempi B. (2005). The organization of recent and remote memories. Nature Rev. Neurosci. 6 119–130. 10.1038/nrn1607 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Frey U., Morris R. G. M. (1998). Synaptic tagging: implications for late maintenance of hippocampal long-term potentiation. Trends Neurosci. 21 181–188. 10.1016/s0166-2236(97)01189-2 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Friedman N. P., Robbins T. W. (2022). The role of prefrontal cortex in cognitive control and executive function. Neuropsychopharmacology 47 72–89. 10.1038/s41386-021-01132-0 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Funahashi S. (2017). Working memory in the prefrontal cortex. Brain Sci. 7 : 49 . 10.3390/brainsci7050049 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gais S., Lucas B., Born J. (2006). Sleep after learning AIDS memory recall. Learn. Mem. 13 259–262. 10.1101/lm.132106 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gallistel C. R., Balsam P. D. (2014). Time to rethink the neural mechanisms of learning and memory. Neurobiol. Learn. Mem. 108 136–144. 10.1016/j.nlm.2013.11.019 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ge Y., Dong Z., Bagot R. C., Howland J. G., Phillips A. G., Wong T. P., et al. (2010). Hippocampal long-term depression is required for the consolidation of Spatial Memory. Proc. Natl Acad. Sci. U.S.A. 107 16697–16702. 10.1073/pnas.1008200107 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gelastopoulos A., Whittington M. A., Kopell N. J. (2019). Parietal low beta rhythm provides a dynamical substrate for a working memory buffer. Proc. Natl Acad. Sci. U.S.A. 116 16613–16620. 10.1073/pnas.1902305116 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Genzel L., Wixted J. T. (2017). “ Cellular and systems consolidation of declarative memory ,” in Cognitive neuroscience of memory consolidation , eds Axmacher N., Rasch B. (Berlin: Springer International Publishing; ), 3–16. 10.1007/978-3-319-45066-7_1 [ CrossRef ] [ Google Scholar ]
  • Ghaleh M., Lacey E. H., Fama M. E., Anbari Z., DeMarco A. T., Turkeltaub P. E. (2019). Dissociable mechanisms of verbal working memory revealed through multivariate lesion mapping. Cereb. Cortex 30 2542–2554. 10.1093/cercor/bhz259 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gilboa A., Moscovitch M. (2021). No consolidation without representation: correspondence between neural and psychological representations in recent and remote memory. Neuron 109 2239–2255. 10.1016/j.neuron.2021.04.025 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Girardeau G., Lopes-dos-Santos V. (2021). Brain neural patterns and the memory function of sleep. Science 374 560–564. 10.1126/science.abi8370 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Goedert K. M., Willingham D. B. (2002). Patterns of interference in sequence learning and prism adaptation inconsistent with the consolidation hypothesis. Learn. Mem. 9 279–292. 10.1101/lm.50102 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Goerke M., Cohrs S., Rodenbeck A., Grittner U., Sommer W., Kunz D. (2013). Declarative memory consolidation during the first night in a sleep lab: the role of REM sleep and Cortisol. Psychoneuroendocrinology 38 1102–1111. 10.1016/j.psyneuen.2012.10.019 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Graves W. W., Grabowski T. J., Mehta S., Gupta P. (2008). The left posterior superior temporal gyrus participates specifically in accessing lexical phonology. J. Cogn. Neurosci. 20 1698–1710. 10.1162/jocn.2008.20113 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hamann S. (2001). Cognitive and neural mechanisms of emotional memory. Trends Cogn. Sci. 5 394–400. 10.1016/S1364-6613(00)01707-1 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Henson R. N., Rugg M. D., Shallice T., Josephs O., Dolan R. J. (1999). Recollection and familiarity in recognition memory: an event-related functional magnetic resonance imaging study. J. Neurosci. 19 3962–3972. 10.1523/jneurosci.19-10-03962.1999 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Herry C., Johansen J. P. (2014). Encoding of fear learning and memory in distributed neuronal circuits. Nat. Neurosci. 17 1644–1654. 10.1038/nn.3869 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hitch G. J., Allen R. J., Baddeley A. D. (2019). Attention and binding in visual working memory: two forms of attention and two kinds of buffer storage. Attent. Percept. Psychophys. 82 280–293. 10.3758/s13414-019-01837-x [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Howard M. D., Skorheim S. W., Pilly P. K. (2022). A model of bi-directional interactions between complementary learning systems for memory consolidation of sequential experiences. Front. Syst. Neurosci. 16 : 972235 . 10.3389/fnsys.2022.972235 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hutchinson J. B., Uncapher M. R., Wagner A. D. (2009). Posterior parietal cortex and episodic retrieval: convergent and divergent effects of attention and memory. Learn. Mem. 16 343–356. 10.1101/lm.919109 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ikkai A., Curtis C. E. (2011). Common neural mechanisms supporting spatial working memory, attention and motor intention. Neuropsychologia 49 1428–1434. 10.1016/j.neuropsychologia.2010.12.020 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ito M., Kuroiwa J., Miyake S. (2003). A neural network model of memory system using hippocampus. Electron. Commun. Japan 86 86–97. 10.1002/ecjc.1010 [ CrossRef ] [ Google Scholar ]
  • Ivanco T. L. (2015). Long-term potentiation and long-term depression. International Encyclopedia of the Social & Behavioral Sciences 14 , 358–365. 10.1016/b978-0-08-097086-8.55034-x [ CrossRef ] [ Google Scholar ]
  • Jerde T. A., Curtis C. E. (2013). Maps of space in human frontoparietal cortex. J. Physiol. 107 510–516. 10.1016/j.jphysparis.2013.04.002 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ji D., Wilson M. A. (2006). Coordinated memory replay in the visual cortex and hippocampus during sleep. Nat. Neurosci. 10 100–107. 10.1038/nn1825 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Josselyn S. A., Köhler S., Frankland P. W. (2015). Finding the engram. Nat. Rev. Neurosci. 16 521–534. 10.1038/nrn4000 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Jung J., Lambon Ralph M. A., Jackson R. L. (2022). Subregions of DLPFC display graded yet distinct structural and functional connectivity. J. Neurosci. 42 3241–3252. 10.1523/jneurosci.1216-21.2022 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kaleem A., Hoessli D. C., Haq I., Walker-Nasir E., Butt A., Iqbal Z., et al. (2011). CREB in long-term potentiation in hippocampus: role of post-translational modifications-studies in silico. J. Cell. Biochem. 112 138–146. 10.1002/jcb.22909 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kamiński J., Rutishauser U. (2019). Between persistently active and activity-silent frameworks: novel vistas on the cellular basis of working memory. Ann. N. Y. Acad. Sci. 1464 64–75. 10.1111/nyas.14213 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kamiński J., Sullivan S., Chung J. M., Ross I. B., Mamelak A. N., Rutishauser U. (2017). Persistently active neurons in human medial frontal and medial temporal lobe support working memory. Nat. Neurosci. 20 590–601. 10.1038/nn.4509 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kandel E. R., Mack S. (2013). Principles of neural science. New York, NY: McGraw-Hill Medical. [ Google Scholar ]
  • Kandel E. R., Dudai Y., Mayford M. R. (2014). The molecular and systems biology of memory. Cell 157 163–186. 10.1016/j.cell.2014.03.001 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Karni A., Meyer G., Rey-Hipolito C., Jezzard P., Adams M. M., Turner R., et al. (1998). The acquisition of skilled motor performance: fast and slow experience-driven changes in primary motor cortex. Proc. Natl. Acad. Sci. U.S.A. 95 861–868. 10.1073/pnas.95.3.861 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kerzel D., Burra N. (2020). Capture by context elements, not attentional suppression of distractors, explains the PD with small search displays. J. Cogn. Neurosci. 32 1170–1183. 10.1162/jocn_a_01535 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kesner R. P. (2017). Memory neurobiology?. Reference Module in Neuroscience and Biobehavioral Psychology 1–12. 10.1016/b978-0-12-809324-5.03089-3 [ CrossRef ] [ Google Scholar ]
  • Kim J., Gulati T., Ganguly K. (2019). Competing roles of slow oscillations and delta waves in memory consolidation versus forgetting. Cell 179 514–526.e13. 10.1016/j.cell.2019.08.040 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Koenigs M., Acheson D. J., Barbey A. K., Solomon J., Postle B. R., Grafman J. (2011). Areas of left perisylvian cortex mediate auditory–verbal short-term memory. Neuropsychologia 49 3612–3619. 10.1016/j.neuropsychologia.2011.09.013 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kosslyn S. M., Thompson W. L., Alpert N. M. (1997). Neural Systems shared by visual imagery and visual perception: a positron emission tomography study. Neuroimage 6 320–334. 10.1006/nimg.1997.0295 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kuhnke P., Kiefer M., Hartwigsen G. (2020). Task-dependent recruitment of modality-specific and multimodal regions during conceptual processing. Cereb. Cortex 30 3938–3959. 10.1093/cercor/bhaa010 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Landmann N., Kuhn M., Piosczyk H., Feige B., Baglioni C., Spiegelhalder K., et al. (2014). The reorganisation of memory during sleep. Sleep Med. Rev. 18 531–541. 10.1016/j.smrv.2014.03.005 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Langille J. J. (2019). Remembering to forget: a dual role for sleep oscillations in memory consolidation and forgetting. Front. Cell. Neurosci. 13 : 71 . 10.3389/fncel.2019.00071 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Langille J. J., Brown R. E. (2018). The synaptic theory of memory: a historical survey and reconciliation of recent opposition. Front. Syst. Neurosci. 12 : 52 . 10.3389/fnsys.2018.00052 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lee H., Chun M. M., Kuhl B. A. (2016). Lower parietal encoding activation is associated with sharper information and Better Memory. Cereb. Cortex 27 , 2486–2499. 10.1093/cercor/bhw097 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Li M., Lu S., Wu Y., Zhong N. (2010). “ Functional segregation of memory encoding and retrieval in the prefrontal cortex ,” in Proceedings of the IEEE/ICME International Conference on Complex Medical Engineering , (Gold Coast, QLD: ), 91–95. 10.1109/iccme.2010.5558865 [ CrossRef ] [ Google Scholar ]
  • Li Q., Song J.-L., Li S.-H., Westover M. B., Zhang R. (2020). Effects of cholinergic neuromodulation on thalamocortical rhythms during NREM sleep: a model study. Front. Comput. Neurosci. 13 : 100 . 10.3389/fncom.2019.00100 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lisman J. E., Grace A. A. (2005). The hippocampal-VTA loop: controlling the entry of information into long-term memory. Neuron 46 703–713. 10.1016/j.neuron.2005.05.002 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Liu J., Zhang H., Yu T., Ren L., Ni D., Yang Q., et al. (2021). Transformative neural representations support long-term episodic memory. Sci. Adv. 7 : eabg9715 . 10.1126/sciadv.abg9715 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Luscher C., Malenka R. C. (2012). NMDA receptor-dependent long-term potentiation and long-term depression (LTP/LTD). Cold Spring Harb. Perspect. Biol. 4 a005710–a005710. 10.1101/cshperspect.a005710 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lynch M. A. (2004). Long-term potentiation and memory. Physiol. Rev. 84 87–136. 10.1152/physrev.00014.2003 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Maass W., Markram H. (2002). Synapses as dynamic memory buffers. Neural Netw. 15 155–161. 10.1016/s0893-6080(01)00144-7 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mackey W. E., Devinsky O., Doyle W. K., Golfinos J. G., Curtis C. E. (2016). Human parietal cortex lesions impact the precision of spatial working memory. J. Neurophysiol. 116 1049–1054. 10.1152/jn.00380.2016 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Malenka R. C., Bear M. F. (2004). LTP and Ltd. Neuron 44 5–21. 10.1016/j.neuron.2004.09.012 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Malenka R. C., Nicoll R. (1999). Long-term potentiation–a decade of progress? Science 285 1870–1874. 10.1126/science.285.5435.1870 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mank M., Griesbeck O. (2008). Genetically encoded calcium indicators. Chem. Rev. 108 1550–1564. 10.1021/cr078213v [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Maquet P. (2001). The role of sleep in learning and memory. Science 294 1048–1052. 10.1126/science.1062856 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Marshall L., Born J. (2007). The contribution of sleep to hippocampus-dependent memory consolidation. Trends Cogn. Sci. 11 442–450. 10.1016/j.tics.2007.09.001 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mayes A. R. (2002). Memory disorders, organic. Encyclopedia of the Human Brain 2 , 759–772. 10.1016/b0-12-227210-2/00199-0 [ CrossRef ] [ Google Scholar ]
  • McGaugh J. L. (2000). Memory–a century of consolidation. Science 287 248–251. 10.1126/science.287.5451.248 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mednick S. C., Cai D. J., Shuman T., Anagnostaras S., Wixted J. T. (2011). An opportunistic theory of cellular and systems consolidation. Trends Neurosci. 34 504–514. 10.1016/j.tins.2011.06.003 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Menon V., D’Esposito M. (2021). The role of PFC networks in cognitive control and executive function. Neuropsychopharmacology 47 90–103. 10.1038/s41386-021-01152-w [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Miyake A., Friedman N. P., Rettinger D. A., Shah P., Hegarty M. (2001). How are visuospatial working memory, executive functioning, and spatial abilities related? A latent-variable analysis. J. Exp. Psychol. 130 621–640. 10.1037/0096-3445.130.4.621 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Moncada D., Ballarini F., Viola H. (2015). Behavioral tagging: a translation of the synaptic tagging and capture hypothesis. Neural Plast. 2015 1–21. 10.1155/2015/650780 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Moro V., Berlucchi G., Lerch J., Tomaiuolo F., Aglioti S. M. (2008). Selective deficit of mental visual imagery with intact primary visual cortex and visual perception. Cortex 44 109–118. 10.1016/j.cortex.2006.06.004 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Moscovitch M., Nadel L., Winocur G., Gilboa A., Rosenbaum R. S. (2006). The cognitive neuroscience of remote episodic, semantic and spatial memory. Curr. Opin. Neurobiol. 16 179–190. 10.1016/j.conb.2006.03.013 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nabavi S., Fox R., Proulx C. D., Lin J. Y., Tsien R. Y., Malinow R. (2014). Engineering A memory with ltd and LTP. Nature 511 348–352. 10.1038/nature13294 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nader K., Hardt O. (2009). A single standard for memory: the case for reconsolidation. Nat. Rev. Neurosci. 10 224–234. 10.1038/nrn2590 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Narayanan N. S., Prabhakaran V., Bunge S. A., Christoff K., Fine E. M., Gabrieli J. D. E. (2005). The role of the prefrontal cortex in the maintenance of verbal working memory: an event-related fMRI analysis. Neuropsychology 19 223–232. 10.1037/0894-4105.19.2.223 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nee D. E., D’Esposito M. (2016). The hierarchical organization of the lateral prefrontal cortex. eLife 5 : e12112 . 10.7554/elife.12112 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ngo H.-V., Fell J., Staresina B. (2020). Sleep spindles mediate hippocampal-neocortical coupling during long-duration ripples. eLife 9 : e57011 . 10.7554/elife.57011 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ober B. A. (2014). Memory, explicit/implicit. Encyclopedia of the Neurological Sciences 2 , 1042–1044. 10.1016/b978-0-12-385157-4.00455-3 [ CrossRef ] [ Google Scholar ]
  • Okuda K., Højgaard K., Privitera L., Bayraktar G., Takeuchi T. (2020). Initial memory consolidation and the synaptic tagging and capture hypothesis. Eur. J. Neurosci. 54 6826–6849. 10.1111/ejn.14902 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Orsini C. A., Maren S. (2012). Neural and cellular mechanisms of fear and extinction memory formation. Neurosci. Biobehav. Rev. 36 1773–1802. 10.1016/j.neubiorev.2011.12.014 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Osaka N., Logie R. H., D’Esposito M. (eds) (2007). The cognitive neuroscience of working memory. Oxford: Oxford University Press. [ Google Scholar ]
  • Panzeri S., Janotte E., Pequeńo-Zurro A., Bonato J., Bartolozzi C. (2023). Constraints on the design of neuromorphic circuits set by the properties of neural population codes. Neuromorphic Computing Eng. 3 : 012001 . 10.1088/2634-4386/acaf9c [ CrossRef ] [ Google Scholar ]
  • Park P., Kang H., Georgiou J., Zhuo M., Kaang B.-K., Collingridge G. L. (2021). Further evidence that CP-ampars are critically involved in synaptic tag and capture at hippocampal CA1 synapses. Mol. Brain 14 : 26 . 10.1186/s13041-021-00737-2 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Park P., Kang H., Sanderson T. M., Bortolotto Z. A., Georgiou J., Zhuo M., et al. (2018). The role of calcium-permeable ampars in long-term potentiation at principal neurons in the rodent hippocampus. Front. Synaptic Neurosci. 10 : 42 . 10.3389/fnsyn.2018.00042 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Patterson K., Nestor P. J., Rogers T. T. (2007). Where do you know what you know? The representation of semantic knowledge in the human brain. Nat. Rev. Neurosci. 8 976–987. 10.1038/nrn2277 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Paulesu E., Frith C. D., Frackowiak R. S. (1993). The neural correlates of the verbal component of working memory. Nature 362 342–345. 10.1038/362342a0 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Perrachione T. K., Ghosh S. S., Ostrovskaya I., Gabrieli J. D., Kovelman I. (2017). Phonological working memory for words and nonwords in cerebral cortex. J. Speech Lang. Hear. Res. 60 1959–1979. 10.1044/2017_jslhr-l-15-0446 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Popa D., Duvarci S., Popescu A. T., Léna C., Paré D. (2010). Coherent amygdalocortical theta promotes fear memory consolidation during paradoxical sleep. Proc. Natl. Acad. Sci. U.S.A. 107 6516–6519. 10.1073/pnas.0913016107 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Postle B. R. (2006). Working memory as an emergent property of the mind and brain. Neuroscience 139 23–38. 10.1016/j.neuroscience.2005.06.005 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Postle B. R. (2015). The cognitive neuroscience of visual short-term memory. Curr. Opin. Behav. Sci. 1 40–46. 10.1016/j.cobeha.2014.08.004 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Puentes-Mestril C., Aton S. J. (2017). Linking network activity to synaptic plasticity during sleep: hypotheses and recent data. Front. Neural Circ. 11 : 61 . 10.3389/fncir.2017.00061 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Qasim S. E., Mohan U. R., Stein J. M., Jacobs J. (2023). Neuronal activity in the human amygdala and hippocampus enhances emotional memory encoding. Nat. Hum. Behav. 7 754–764. 10.1038/s41562-022-01502-8 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Quentin R., King J.-R., Sallard E., Fishman N., Thompson R., Buch E. R., et al. (2019). Differential brain mechanisms of selection and maintenance of information during working memory. J. Neurosci. 39 3728–3740. 10.1523/jneurosci.2764-18.2019 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Rakowska M., Bagrowska P., Lazari A., Navarrete M., Abdellahi M. E., Johansen-Berg H., et al. (2022). Cueing motor memory reactivation during NREM sleep engenders learning-related changes in precuneus and sensorimotor structures . bioRxiv 2022.01.-27.477838. 10.1101/2022.01.27.477838 [ CrossRef ] [ Google Scholar ]
  • Rasch B., Born J. (2007). Maintaining memories by reactivation. Curr. Opin. Neurobiol. 17 698–703. 10.1016/j.conb.2007.11.007 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Rasch B., Born J. (2013). About sleep’s role in memory. Physiol. Rev. 93 681–766. 10.1152/physrev.00032.2012 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ratcliffe O., Shapiro K., Staresina B. P. (2022). Fronto-medial theta coordinates posterior maintenance of working memory content. Curr. Biol. 32 2121–2129.e3. 10.1016/j.cub.2022.03.045 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Reber P. J. (2008). Cognitive neuroscience of declarative and nondeclarative memory. Hum. Learn. Biol. Brain Neurosci. 139 113–123. 10.1016/s0166-4115(08)10010-3 [ CrossRef ] [ Google Scholar ]
  • Redondo R. L., Morris R. G. M. (2011). Making memories last: the synaptic tagging and capture hypothesis. Nat. Rev. Neurosci. 12 17–30. 10.1038/nrn2963 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ren Z., Zhang Y., He H., Feng Q., Bi T., Qiu J. (2019). The different brain mechanisms of object and spatial working memory: voxel-based morphometry and resting-state functional connectivity. Front. Hum. Neurosci. 13 : 248 . 10.3389/fnhum.2019.00248 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Repovs G., Baddeley A. (2006). The multi-component model of working memory: explorations in experimental cognitive psychology. Neuroscience 139 5–21. 10.1016/j.neuroscience.2005.12.061 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Reyes-Resina I., Samer S., Kreutz M. R., Oelschlegel A. M. (2021). Molecular mechanisms of memory consolidation that operate during sleep. Front. Mol. Neurosci. 14 : 767384 . 10.3389/fnmol.2021.767384 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ribeiro S. (2007). Novel experience induces persistent sleep-dependent plasticity in the cortex but not in the hippocampus. Front. Neurosci. 1 :43–55. 10.3389/neuro.01.1.1.003.2007 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Richter-Levin G., Akirav I. (2003). Emotional tagging of memory formation—In the search for neural mechanisms. Brain Res. Rev. 43 247–256. 10.1016/j.brainresrev.2003.08.005 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ritchey M., Dolcos F., Cabeza R. (2008). Role of amygdala connectivity in the persistence of emotional memories over time: an event-related fmri investigation. Cereb. Cortex 18 2494–2504. 10.1093/cercor/bhm262 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Robin J., Moscovitch M. (2017). Details, gist and schema: hippocampal–neocortical interactions underlying recent and remote episodic and spatial memory. Curr. Opin. Behav. Sci. 17 114–123. 10.1016/j.cobeha.2017.07.016 [ CrossRef ] [ Google Scholar ]
  • Robins S. K. (2020). Stable engrams and neural dynamics. Philos. Sci. 87 1130–1139. 10.1086/710624 [ CrossRef ] [ Google Scholar ]
  • Roediger H. L., Butler A. C. (2011). The critical role of retrieval practice in long-term retention. Trends Cogn. Sci. 15 20–27. 10.1016/j.tics.2010.09.003 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Roldan-Valadez E., García-Lázaro H., Lara-Romero R., Ramirez-Carmona R. (2012). Neuroanatomy of episodic and semantic memory in humans: a brief review of neuroimaging studies. Neurol. India 60 : 613 . 10.4103/0028-3886.105196 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Saffran N. M. (1997). Language and auditory-verbal short-term memory impairments: evidence for common underlying processes. Cogn. Neuropsychol. 14 641–682. 10.1080/026432997381402 [ CrossRef ] [ Google Scholar ]
  • Salvato G., Peviani V., Scarpa P., Francione S., Castana L., Gallace A., et al. (2021). Investigating visuo-spatial neglect and visual extinction during intracranial electrical stimulations: the role of the right inferior parietal cortex. Neuropsychologia 162 : 108049 . 10.1016/j.neuropsychologia.2021.108049 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sarnthein J., Petsche H., Rappelsberger P., Shaw G. L., von Stein A. (1998). Synchronization between prefrontal and posterior association cortex during human working memory. Proc. Natl. Acad. Sci. U.S.A. 95 7092–7096. 10.1073/pnas.95.12.7092 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Schiller D., Monfils M.-H., Raio C. M., Johnson D. C., LeDoux J. E., Phelps E. A. (2009). Preventing the return of fear in humans using reconsolidation update mechanisms. Nature 463 49–53. 10.1038/nature08637 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Schlichting M. L., Preston A. R. (2015). Memory integration: neural mechanisms and implications for behavior. Curr. Opin. Behav. Sci. 1 1–8. 10.1016/j.cobeha.2014 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sellers K. K., Yu C., Zhou Z. C., Stitt I., Li Y., Radtke-Schuller S., et al. (2016). Oscillatory dynamics in the frontoparietal attention network during sustained attention in the ferret. Cell Rep. 16 2864–2874. 10.1016/j.celrep.2016.08.055 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sendi M. S., Kanta V., Inman C. S., Manns J. R., Hamann S., Gross R. E., et al. (2020). “ Amygdala stimulation leads to functional network connectivity state transitions in the hippocampus ,” in Proceedings of the 2020 42nd Annual International Conference of the IEEE Engineering in Medicine & Biology Society (EMBC) , (Montreal, QC: ). 10.1109/embc44109.2020.9176742 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Serences J. T. (2016). Neural mechanisms of information storage in visual short-term memory. Vis. Res. 128 53–67. 10.1016/j.visres.2016.09.010 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Shafritz K. M., Gore J. C., Marois R. (2002). The role of the parietal cortex in visual feature binding. Proc. Natl. Acad. Sci. U.S.A. 99 10917–10922. 10.1073/pnas.152694799 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Shallice T. (2002). “ Fractionation of the supervisory system ,” in Principles of frontal lobe function , eds Stuss D. T., Knight R. T. (Oxford: Oxford University Press; ), 261–277. 10.1093/acprof:oso/9780195134971.003.0017 [ CrossRef ] [ Google Scholar ]
  • Siegel J. M. (2001). The REM sleep-memory consolidation hypothesis. Science 294 1058–1063. 10.1126/science.1063049 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Smith E. E., Jonides J., Marshuetz C., Koeppe R. A. (1998). Components of verbal working memory: evidence from neuroimaging. Proc. Natl. Acad. Sci. U.S.A. 95 876–882. 10.1073/pnas.95.3.876 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sobczak J. (2017). Mechanisms of memory consolidation (dissertation). PhD thesis. York: University of York. [ Google Scholar ]
  • Sorensen K. E. (2009). The connections of the hippocampal region new observations on efferent connections in the guinea pig, and their functional implications*. Acta Neurol. Scand. 72 550–560. 10.1111/j.1600-0404.1985.tb00914.x [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Speranza L., di Porzio U., Viggiano D., de Donato A., Volpicelli F. (2021). Dopamine: the neuromodulator of long-term synaptic plasticity, reward and movement control. Cells 10 : 735 . 10.3390/cells10040735 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Squire L. R., Wixted J. T. (2011). The cognitive neuroscience of human memory since H.M. Annu. Rev. Neurosci. 34 259–288. 10.1146/annurev-neuro-061010-113720 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Squire L. R., Zola S. M. (1996). Structure and function of declarative and nondeclarative memory systems. Proc. Natl. Acad. Sci. U.S.A. 93 13515–13522. 10.1073/pnas.93.24.13515 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Squire L. R., Genzel L., Wixted J. T., Morris R. G. (2015). Memory consolidation. Cold Spring Harb. Perspect. Biol. 7 : a021766 . 10.1101/cshperspect.a021766 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Squire L. R., Stark C. E. L., Clark R. E. (2004). The medial temporal lobe. Annu. Rev. Neurosci. 27 279–306. 10.1146/annurev.neuro.27.070203.144130 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Sreenivasan K. K., Curtis C. E., D’Esposito M. (2014). Revisiting the role of persistent neural activity during working memory. Trends Cogn. Sci. 18 82–89. 10.1016/j.tics.2013.12.001 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Stickgold R. (2005). Sleep-dependent memory consolidation. Nature 437 1272–1278. 10.1038/nature04286 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Swanson R. A., Levenstein D., McClain K., Tingley D., Buzsáki G. (2020). Variable specificity of memory trace reactivation during hippocampal sharp wave ripples. Curr. Opin. Behav. Sci. 32 126–135. 10.1016/j.cobeha.2020.02.008 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Takehara-Nishiuchi K. (2020). Prefrontal–hippocampal interaction during the encoding of New Memories. Brain Neurosci. Adv. 4 239821282092558 . 10.1177/2398212820925580 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Takeuchi T., Duszkiewicz A. J., Morris R. G. (2014). The synaptic plasticity and memory hypothesis: encoding, storage and persistence. Philos. Trans. R. Soc. B Biol. Sci. 369 : 20130288 . 10.1098/rstb.2013.0288 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tamminen J., Payne J. D., Stickgold R., Wamsley E. J., Gaskell M. G. (2010). Sleep spindle activity is associated with the integration of new memories and existing knowledge. The Journal of Neuroscience 30 14356–14360. 10.1523/jneurosci.3028-10.2010 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Thomas Yeo B. T., Krienen F. M., Sepulcre J., Sabuncu M. R., Lashkari D., Hollinshead M., et al. (2011). The organization of the human cerebral cortex estimated by intrinsic functional connectivity. J. Neurophysiol. 106 1125–1165. 10.1152/jn.00338.2011 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Thompson R. F., Kim J. J. (1996). Memory systems in the brain and localization of a memory. Proc. Natl. Acad. Sci. U.S.A. 93 13438–13444. 10.1073/pnas.93.24.13438 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Thuault S. J., Malleret G., Constantinople C. M., Nicholls R., Chen I., Zhu J., et al. (2013). Prefrontal cortex HCN1 channels enable intrinsic persistent neural firing and executive memory function. J. Neurosci. 33 13583–13599. 10.1523/jneurosci.2427-12.2013 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Todd J. J., Marois R. (2005). Posterior parietal cortex activity predicts individual differences in visual short-term memory capacity. Cogn. Affect. Behav. Neurosci. 5 144–155. 10.3758/cabn.5.2.144 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tononi G., Cirelli C. (2003). Sleep and synaptic homeostasis: a hypothesis. Brain Res. Bull. 62 143–150. 10.1016/j.brainresbull.2003.09.004 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tononi G., Cirelli C. (2014). Sleep and the price of plasticity: from synaptic and cellular homeostasis to memory consolidation and integration. Neuron 81 12–34. 10.1016/j.neuron.2013.12.025 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tulving E. (1985). Memory and consciousness. Can. Psychol. 26 1–12. 10.1037/h0080017 [ CrossRef ] [ Google Scholar ]
  • van Kesteren M. T. R., Ruiter D. J., Fernández G., Henson R. N. (2012). How schema and novelty augment memory formation. Trends Neurosci. 35 211–219. 10.1016/j.tins.2012.02.001 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • van Kesteren M. T., Brown T. I., Wagner A. D. (2016). Interactions between memory and new learning: insights from fMRI multivoxel pattern analysis. Front. Syst. Neurosci. 10 : 46 . 10.3389/fnsys.2016.00046 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Vilberg K. L., Rugg M. D. (2008). Memory retrieval and the PARIETAL CORTEX: a review of evidence from a dual-process perspective. Neuropsychologia 46 1787–1799. 10.1016/j.neuropsychologia.2008.01.004 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Vishnoi S., Naseem M., Raisuddin S., Parvez S. (2018). Behavioral tagging: plausible involvement of PKMZ, ARC and role of neurotransmitter receptor systems. Neurosci. Biobehav. Rev. 94 210–218. 10.1016/j.neubiorev.2018.07.009 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wang J. H., Wang D., Gao Z., Chen N., Lei Z., Cui S., et al. (2016). Both glutamatergic and Gabaergic neurons are recruited to be associative memory cells. Biophys. J. 110 : 481a . 10.1016/j.bpj.2015.11.2571 [ CrossRef ] [ Google Scholar ]
  • Wang J.-H., Cui S. (2018). Associative memory cells and their working principle in the brain. F1000Research 7 : 108 . 10.12688/f1000research.13665.1 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wig G. S., Buckner R. L., Schacter D. L. (2009). Repetition priming influences distinct brain systems: evidence from task-evoked data and resting-state correlations. J. Neurophysiol. 101 2632–2648. 10.1152/jn.91213.2008 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Williams J. N. (2020). The neuroscience of implicit learning. Language Learning 70 255–307. 10.1111/lang.12405 [ CrossRef ] [ Google Scholar ]
  • Wiltgen B. J., Zhou M., Cai Y., Balaji J., Karlsson M. G., Parivash S. N., et al. (2010). The hippocampus plays a selective role in the retrieval of detailed contextual memories. Curr. Biol. 20 1336–1344. 10.1016/j.cub.2010.06.068 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Winocur G., Moscovitch M. (2011). Memory transformation and systems consolidation. J. Int. Neuropsychol. Soc. 17 766–780. 10.1017/S1355617711000683 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wolf R. C., Vasic N., Walter H. (2006). Differential activation of ventrolateral prefrontal cortex during working memory retrieval. Neuropsychologia 44 2558–2563. 10.1016/j.neuropsychologia.2006.05.015 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Xu Y., Chun M. M. (2006). Dissociable neural mechanisms supporting visual short-term memory for objects. Nature 440 91–95. 10.1038/nature04262 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Xu Y., Zou P., Cohen A. E. (2017). Voltage imaging with genetically encoded indicators. Curr. Opin. Chem. Biol. 39 1–10. 10.1016/j.cbpa.2017.04.005 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Yang P., Wang M., Luo C., Ni X., Li L. (2022). Dissociable causal roles of the frontal and parietal cortices in the effect of object location on object identity detection: a TMS study. Exp. Brain Res. 240 1445–1457. 10.1007/s00221-022-06344-4 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Yu R., Han B., Wu X., Wei G., Zhang J., Ding M., et al. (2023). Dual-functional network regulation underlies the Central Executive System in working memory. Neuroscience 524 158–180. 10.1016/j.neuroscience.2023.05.025 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Yu S., Rempel S., Gholamipourbarogh N., Beste C. (2022). A ventral stream-prefrontal cortex processing cascade enables working memory gating dynamics. Commun. Biol. 5 : 1086 . 10.1038/s42003-022-04048-7 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Zeithamova D., Preston A. R. (2010). Flexible memories: differential roles for medial temporal lobe and prefrontal cortex in cross-episode binding. J. Neurosci. 30 14676–14684. 10.1523/jneurosci.3250-10.2010 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Zhu Y., Zeng Y., Ren J., Zhang L., Chen C., Fernandez G., et al. (2022). Emotional learning retroactively promotes memory integration through rapid neural reactivation and reorganization. eLife 11 : e60190 . 10.7554/elife.60190 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Zielinski M. C., Tang W., Jadhav S. P. (2018). The role of replay and Theta sequences in mediating hippocampal-prefrontal interactions for memory and cognition. Hippocampus 30 60–72. 10.1002/hipo.22821 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Zlotnik G., Vansintjan A. (2019). Memory: an extended definition. Front. Psychol. 10 : 2523 . 10.3389/fpsyg.2019.02523 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Reviews / Why join our community?
  • For companies
  • Frequently asked questions

presentation on human memory

The Properties of Human Memory and Their Importance for Information Visualization

It is important to know that while neuroscience has progressed dramatically over the last decades; there is no complete understanding of how human memory works. We know, for example, that data in the brain is stored in clusters of neurons but we don’t know how, precisely, it is stored or even how it is encoded. Thus when it comes to understanding memory from a design perspective we will examine certain properties of human memory that are commonly understood to be correct.

Human memory doesn’t exist in isolation; the brain isn’t just responsible for memorizing things but also for processing the data and acting on that data. Much of our memory and much of the information we receive is visual and it is with visual memories that the designer is mainly concerned.

Three Types of Memory

There are three main types of memory that are processed in the brain:

Sensory Memories

Short-term Memories

Long-term Memories

presentation on human memory

Author/Copyright holder: JSpudeman. Copyright terms and licence: Public Domain.

Sensory memories are the memories which are stored for tiny time periods and which originate from our sensory organs (such as our eyes or our nose). They are typically retained for less than 500 milliseconds.

Visual sensory memory is often known as iconic memory. Sensory visual memories are the raw information that the brain receives (via the optic nerve) from the eye. We store and process sensory memories automatically – that is without any conscious effort to do so.

The processing of this information is called preattentive processing (e.g. it happens prior to our paying attention to the information). It is a limited form of processing which does not attempt to make sense of the whole image received but rather to a small set of features of the image – such as colors , shapes, tilt, curvature, contrast, etc.

presentation on human memory

Author/Copyright holder: Was a bee. Copyright terms and licence: CC BY-SA 2.5

It is sensory memory which draws your attention to the strawberries in this graphic.

Short-Term Memories

Short-term memory is used to process sensory memories which are of interest to us – for whatever reason. The sensory memory is transferred to the short-term memory where it may be processed for up to a minute (though if the memory is rehearsed – e.g. repeated – it may remain in short-term memory for a longer period up to a few hours in length).

Short-term memory is of limited capacity. Experiments conducted by, among others, George A Miller the psychologist, and reported in his paper “The Magical Number Seven, plus or minus two” suggest that we can store between 5 and 9 similar items in short-term memory at the most.

This capacity can be increased by a process known as “ chunking ”. This is where we group items to form larger items. So, for example, you can memorize a 12 digit phone number in short-term memory by taking digits in pairs (35) rather than singly (3 and 5) which gives you 6 chunks to remember (which falls between 5 and 9) rather than 12 digits (which exceeds the capacity of short-term memory).

Chunking can occur visually as well as through combination of numeric or alpha-numeric attributes. A common example of this would be in a bar chart where a single bar may represent a chunk of information.

This is useful to the visual designer because it allows a visual representation of information to be easily processed in short-term memory and for that representation to offer more complex insights than an initial examination of the capacity of short-term memory might allow.

Author/Copyright holder: Michaelchilliard. Copyright terms and licence: CC BY-SA 3.0

This graph, above, shows how information recall is limited from the short term memory and recall becomes worse when asked to recall a sequence backwards.

Long-Term Memories

In most instances the memories transferred to our short-term memories are quickly forgotten. This is, probably, a good thing. If we didn’t forget the huge volumes of information that we perceive on a daily basis we could well become overloaded with information and find processing it in a meaningful way soon became impossible.

In order for most memories to transfer from short-term to long-term memory – conscious effort must be made to effect the transfer. This is why students review for examinations; the repeated application of information or rehearsing of information enables the transfer of the material they are studying to long-term memory.

presentation on human memory

Author/Copyright holder: Omphalosskeptic. Copyright terms and licence: CC BY-SA 3.0

It is also possible for a long-term memory to evolve through a meaningful association in the brain. For example, we know that a static shock is painful even if we are only shocked once. It doesn’t take repeated shocks to memorize that. The meaningful connection between the pain and the shock allow us to process the memory long-term. In fact strong emotional or physical connections are often the easiest way for something to enter long-term memory.

The image above is of a Van de Graaf Generator which can be used to generate static electricity – you can then touch the generator and another person to give them a static shock. It’s worth remembering that they won’t come back for a 2nd attempt…

It is worth noting that majority of designs and in particular, information visualizations, will not be committed to long-term memory. It may be that the conclusions or understanding they bring will be transferred to long-term memory (usually through revision or application) but the design itself will not.

The vast majority of interaction between the user and an information visualization will occur in sensory and short-term memory.

The Take Away

The key link between design (and in particular information visualization design) and human memory is that interaction takes place in sensory and short-term memory for most users. This means paying careful attention to not providing more than 9 chunks of data in a visualization (and ideally no more than 5) and trying to ensure that you use a single visualization to convey the information because once someone’s attention moves from one image to another – the first is quickly forgotten.

Edward Tufte, the world’s leading authority in information visualization asks; “Can the same image prompt different stories and memories in different people? That’s a good test for a “super-graphic.” By better understanding memory, perhaps we can create super graphics more easily.

References & Where to Learn More:

Course: Information Visualization

George A. Miller. The magical number seven, plus or minus two. The Psychological Review , 63(2):81–97, 1956.

Hero Image: Author/Copyright holder: Henry Vandyke Carter. Copyright terms and licence: Public Domain.

Information Visualization

presentation on human memory

Get Weekly Design Tips

Topics in this article, what you should read next, information overload, why it matters and how to combat it.

presentation on human memory

  • 1.1k shares
  • 3 years ago

Recalling Color Theory Keywords: a way to refresh your memories!

presentation on human memory

How to Design an Information Visualization

presentation on human memory

How to Visualize Your Qualitative User Research Results for Maximum Impact

presentation on human memory

Preattentive Visual Properties and How to Use Them in Information Visualization

presentation on human memory

  • 5 years ago

How to Conduct Focus Groups

presentation on human memory

Information Visualization – A Brief Introduction

presentation on human memory

Visual Mapping – The Elements of Information Visualization

presentation on human memory

Guidelines for Good Visual Information Representations

presentation on human memory

  • 4 years ago

How to Show Hierarchical Data with Information Visualization

presentation on human memory

Open Access—Link to us!

We believe in Open Access and the  democratization of knowledge . Unfortunately, world-class educational materials such as this page are normally hidden behind paywalls or in expensive textbooks.

If you want this to change , cite this article , link to us, or join us to help us democratize design knowledge !

Privacy Settings

Our digital services use necessary tracking technologies, including third-party cookies, for security, functionality, and to uphold user rights. Optional cookies offer enhanced features, and analytics.

Experience the full potential of our site that remembers your preferences and supports secure sign-in.

Governs the storage of data necessary for maintaining website security, user authentication, and fraud prevention mechanisms.

Enhanced Functionality

Saves your settings and preferences, like your location, for a more personalized experience.

Referral Program

We use cookies to enable our referral program, giving you and your friends discounts.

Error Reporting

We share user ID with Bugsnag and NewRelic to help us track errors and fix issues.

Optimize your experience by allowing us to monitor site usage. You’ll enjoy a smoother, more personalized journey without compromising your privacy.

Analytics Storage

Collects anonymous data on how you navigate and interact, helping us make informed improvements.

Differentiates real visitors from automated bots, ensuring accurate usage data and improving your website experience.

Lets us tailor your digital ads to match your interests, making them more relevant and useful to you.

Advertising Storage

Stores information for better-targeted advertising, enhancing your online ad experience.

Personalization Storage

Permits storing data to personalize content and ads across Google services based on user behavior, enhancing overall user experience.

Advertising Personalization

Allows for content and ad personalization across Google services based on user behavior. This consent enhances user experiences.

Enables personalizing ads based on user data and interactions, allowing for more relevant advertising experiences across Google services.

Receive more relevant advertisements by sharing your interests and behavior with our trusted advertising partners.

Enables better ad targeting and measurement on Meta platforms, making ads you see more relevant.

Allows for improved ad effectiveness and measurement through Meta’s Conversions API, ensuring privacy-compliant data sharing.

LinkedIn Insights

Tracks conversions, retargeting, and web analytics for LinkedIn ad campaigns, enhancing ad relevance and performance.

LinkedIn CAPI

Enhances LinkedIn advertising through server-side event tracking, offering more accurate measurement and personalization.

Google Ads Tag

Tracks ad performance and user engagement, helping deliver ads that are most useful to you.

Share Knowledge, Get Respect!

or copy link

Cite according to academic standards

Simply copy and paste the text below into your bibliographic reference list, onto your blog, or anywhere else. You can also just hyperlink to this article.

New to UX Design? We’re giving you a free ebook!

The Basics of User Experience Design

Download our free ebook The Basics of User Experience Design to learn about core concepts of UX design.

In 9 chapters, we’ll cover: conducting user interviews, design thinking, interaction design, mobile UX design, usability, UX research, and many more!

New to UX Design? We’re Giving You a Free ebook!

Jeremy Gutsche Innovation Keynote Speaker

15 Presentations on the Human Memory

presentation on human memory

From The Art of Lists to Gaming and the Brain

Effects of Exercise on the Brain

  • Announcements
  • Copilot Studio

Microsoft Copilot Studio: Building copilots with agent capabilities

presentation on human memory

Omar Aftab , Vice President, Conversational AI , Tuesday, May 21, 2024

Facebook

At Microsoft Build 2024 , we’re excited to announce a host of new powerful capabilities in   Microsoft Copilot Studio —t he single conversational AI tool you can use to create your very own custom copilots or extend Microsoft C opilot experiences with your own enterprise data and scenarios.

The first of these are c opilots that can now act as independent agents— ones that can be triggered by events— not just conversation— and can automa te and orchestrate complex, long-running business processes with more autonomy and less human intervention.

For instance, consider the potential of a copilot that can react when an email arrives, look up the sender’s details, see their previous communications, and use generative AI to trigger the appropriate chain of actions in their response. From understanding the intent of the email, to look ing up the sender’s details and account , see ing their previous communications, checking inventory,   responding to the sender asking for their preferences, and then taking the appropriate actions to close a ticket — orchestrating and shepherding an entire process over days.  

With such capabilities, copilots are evolving from those that work with you to those that work for you. They can be designed to handle specific roles or functions, such as IT, marketing, sales, customer success, and finance across various industries, including travel, retail, and financial services.  

With these new capabilities, here are some examples of the kinds of copilots our customers can build:  

  • IT help desk .  IT support is complex, involving tickets, order numbers, approvals, and stock levels . O pening and closing a ticket can be a long-running task that spans days. A copilot can now handle this process, interfacing with IT service management applications, resolving IT tickets with context and memory, creating purchase orders for device refresh, and reaching out and getting managers approvals — all independently .
  • Employee onboarding . Onboarding new employees is often expensive and slow. Now, imagine you’re a new hire. A copilot greets you, reasons over HR data, and answers your questions. It introduces you to your buddy, provides training and deadlines, assists with forms, and sets up your first week of meetings. Throughout all of this, the copilot is in touch, guiding you through the weeks -long onboarding and account set up processes.  
  • Personal concierge for sales and service . Balancing exceptional customer experience while meeting ambitious revenue goals can be challenging. When a copilot serves guests, i t can use the memory of previous conversations with guests to remember their preferences, make reservations, handle complaints, and answer questions related to the products and services on offer. The copilot learns from its interactions and proposes new ways of handling customer scenarios. By doing so, copilots can increase upsell and attachment rates, driving revenue for the resort while simultaneously enhancing guest experience, satisfaction rates, and repeat business.

Let’s dig deeper into a few of the underlying capabilities that make all this possible:

  • Asynchronous orchestration of complex tasks . The first is the ability to use generative AI- powered   planning and reasoning to manage complex, multi step, long-running tasks. For example, reacting to a new order means determining the need to verify inventory, trigger ing the right payment processes, pinging a supervisor for approval if the amount is above a certain threshold, and replying with a confirmation. Many of these events can take hours—or even days— to complete, but the copilot will run through them , maintaining the necessary state and context to do so.
  • Memory and context . One of the frustrating things about support has traditionally been having to repeat information: who you are, what your policy number is, what your address is. There is no continuity of conversation. Copilots will now learn from previous conversations from the users and utilize this knowledge to continually personalize interactions . A copilot may not need to ask you for your laptop model or your address when you call again for the same issue. Conversations will thus become long-running, contextual, and deeply personalized.
  • Monitor, learn, and improve . Copilots can now learn and adapt, offering monitoring and teaching capabilities to make their interactions better. Each copilot records a comprehensive history of its activities, providing transparency into its performance, including user interactions, actions taken, and feedback received, and you can see what decisions it made — and correct and teach them — with just a few clicks.

Screenshot of the in-product experience for training copilots with agent capabilities in Microsoft Copilot Studio

  • Delegation with confidence and guardrails . When developing copilots with agent capabilities, establishing clear boundaries is paramount. Copilots operate strictly within the confines of the maker-defined instructions, knowledge, and actions. The data sources linked to the copilot adhere to stringent security measures and controls, managed through the unified admin center of Copilot Studio. This includes data loss prevention, robust authentication protocols, and more.

The se advanced new capabilities in Copilot Studio are currently accessible to customers participating in a limited private preview  where organizations such as Centro de la Familia are excited to explore agent capabilities that support teachers and case workers, allowing them to spend less time on administrative tasks and more time working with children, ultimately leading to better child outcomes . Based on feedback from program participants, we will continue to iterate and refine these capabilities for broader access in a preview planned for later this year .  

Additional innovations with Copilot Studio

There’s a lot more to share at Microsoft Build with Copilot Studio, and we’ll touch on just a few of our new capabilities here. To learn more — just sign up and try it out for yourself here .

It’s easier than ever to create c opilots .  With Copilot Studio, creating and testing copilots is now incredibly simple. You can create your copilot with our brand new conversationally driven experience — simply describe what you want it to do, and what knowledge you want it to have, and Copilot Studio will create your very own c opilot. You can then immediately test it out, add additional capabilities, such as your own actions, APIs, and enterprise knowledge — and then publish it live with a few clicks.

Screenshot of the homepage of Microsoft Copilot Studio

Connect all your enterprise data with Copilot c onnectors .   Customers want copilots connected with data from their own enterprises business systems and apps. Copilot connectors enable anyone to ground their copilot in business and collaboration data. This makes it possible for copilots to use various data sources, including public websites, SharePoint, OneDrive, Microsoft Dataverse tables, Microsoft Fabric OneLake (coming this calendar year), Microsoft Graph, as well as leading third-party apps. You can even create your own custom generative prompts to configure how a copilot handles a response from an API or connector.

Screenshot of the available knowledge sources in Microsoft Copilot Studio

Here are a few examples of how Copilot connectors can transform copilot experiences for specific personas or functions:

  • Legal and Compliance . Navigate complex legal landscapes with a Copilot extension that queries specific legal datasets, ensuring controlled and compliant responses without overwhelming users with extraneous information.
  • HR Helper . Assist employees with accessing essential resources for benefits and PTO policies, and even book time off directly through Copilot.
  • Incident Report Coordinator . Workers can locate the right documentation, report incidents, and track them efficiently, all within the context of the chat.

Starting in June 2024, developers can access the preview for Copilot connectors and stay informed on updates here .

Conversational analytics (private preview) : One of the most common asks from customers has been the need for deeper insight into what their copilot is doing, how generative AI is responding, when it was unable to give the right answers and why — and recommendations on what to do to improve it.

Screenshot of the conversational analytics experience in Microsoft Copilot Studio

Templates : If simply describing your copilot to build it wasn’t easy enough, Copilot Studio will now also include a variety of pre-built copilot samples for departments and industries. Some templates — such as Safe Travels for comprehensive travel support, Organization Navigator for organizational clarity, Kudos Copilot for fostering recognition, Wellness for employee health insights — are available now, with many more releasing in the coming months.

Enhanced security and controls (public preview ) : Administrators can now configure advanced settings beyond the default security measures and controls. With Microsoft Purview , Copilot Studio administrators gain access to more detailed governance tools, including audit logs, inventory capabilities, and sensitivity labels. They will be able to review comprehensive audit logs that cover tenant-wide usage, inventory (with API support), and tenant hygiene (such as data loss prevention violations and inactive copilots), enabling them to effectively monitor business impact. Both creators and end-users will be able to view sensitivity labels when responses are generated using AI-powered answers based on SharePoint documents.

With all the amazing innovations, numerous organizations are using Copilot Studio to build transformative generative AI-powered solutions. Check out this story from Nsure on how they are using Copilot Studio:

Get started today with Copilot Studio

This is just a glimpse of all the exciting innovation around copilots and Copilot Studio — we have a host of exciting new capabilities to share in our sessions at Build. So, join us in watching the sessions below, and try out Copilot Studio yourself and build and share your very own copilot in minutes.

Watch the sessions at Microsoft Build:

  • “ Microsoft Build opening keynote ”
  • “ Next generation AI for developers with the Microsoft Cloud ”
  • “ Shaping next-gen development: the future of Copilot in Power Platform ”

Deeper dives:

  • Breakout: “ What’s new with Microsoft Copilot Studio ”
  • Breakout with demos: “ Build your own copilot with Microsoft Copilot Studio ”
  • Breakout with demos: “ Build Microsoft Copilot extensions with Copilot Studio ”
  • Demo (live only): “ Build your own Copilot extension with Microsoft Copilot Studio ”

human memory

Human Memory

Jan 04, 2020

200 likes | 244 Views

Human Memory. Sensory memory Short term memory Long term memory. What is memory?. It is the ability to store and retrieve the information Much of our everyday activity relies on memory We need to understand some of the capabilities and limitations of human memory to answer these

Share Presentation

  • sensory memory
  • echoic memory
  • iconic memory
  • sensory memory exists
  • echoic memory acoustic stimulus

lmallory

Presentation Transcript

Human Memory Sensory memory Short term memory Long term memory

What is memory? • It is the ability to store and retrieve the information • Much of our everyday activity relies on memory • We need to understand some of the capabilities and limitations of human memory to answer these • How does memory works? • How do we remember some arbitrary list ? • Why do some people remember more easily than others?

Attention Rehearsal Memory There are three types of memory function: Sensory memories Short-term memory or working memory Long-term memory Selection of stimuli governed by level of arousal.

Sensory memory • It act as “buffers” for stimuli received through senses • A sensory memory exists for each sensory channel • iconic memory: visual stimulus • echoic memory: acoustic stimulus • haptic memory: touch stimulus • Continuously overwritten by new information coming in on these channels

Sensory memory • It act as “buffers” for stimuli received through senses • A sensory memory exists for each sensory channel • iconic memory: visual stimulus • echoic memory: acoustic stimulus • haptic memory: touch stimulus • Continuously overwritten by new information coming in on these channels • Examples • Iconic memory: At firework displays where moving “sparklers” leave a persistence image, 0.5 sec • Echoic memory: brief “play back” of information

Sensory memory • Information is passed from sensory memory into short-term memory by attention • Attention is the concentration of the mind on one out of a number of competing stimuli or thoughts • It is clear we are able to focus our attention to one thing at a time • This is due to the limited capacity of our sensory memory • Otherwise overloaded

Sensory memory • We can choose which stimuli to attend to, and this choice is our level of interest or need • This explains the “cocktail party phenomenon” • “We can attend to one conversation over the background noise, but we may choose to switch our attention to a conversation across the room if we here our name mentioned” • Information received by sensory memories is quickly passed into more permanent or overwritten and lost

Short-term memory (STM) • It act as a “Scratch-pad” for temporary recall of information • Examples: • Calculate the multiplication 35x6 in your head • Comprehensive test • rapid access ~ 70ms • rapid decay ~ 200ms

Short-term memory (STM) • STM also has limited capacity- 7± 2 chunks • There are two methods for measuring memory capacity • Recall the sequence in order • Recall the sequence in any order

Try this! 212348278493202

Try this! 212348278493202 01 21 414 2626

Short-term memory (STM) • The successful formation of a chunk is known as “closure” • In Design Focus: Cashing in • ATM machine

Short Term Memory (STM): • In design Focus: 7± 2 revisited • List, menu and groups of items should be no more than 7 items long • Command line interfaces

Short-term memory (STM) • Patterns can be useful as aids to memory

Try this! • Patterns can be useful as aids to memory • HEC ATR ANU PTH ETR EET

Try this! • Patterns can be useful as aids to memory • HEC ATR ANU PTH ETR EET • The sequence is easy to recall (the cat ran up the tree)

Short-term memory (STM):The “recency effect” • Recall of the last words presented is better than recall of those in the middle. • This is known as the “recency effect”

  • More by User

Chapter 7: Human Memory

Chapter 7: Human Memory

Chapter 7: Human Memory. The office. Human Memory: Basic Questions. How does information get into memory? How is information maintained in memory? How is information pulled back out of memory?. Figure 7.2 Three key processes in memory. Encoding: Getting Information Into Memory.

949 views • 34 slides

Human Memory

Human Memory. Don’t ask too much of mere mortals Material mainly from Dix et al chapter 1. Learning outcomes. Describe the major categories of human memory Describe the major organization structures of long term memory How are these organization structures reflected in UI design. Attention.

675 views • 21 slides

Human Memory

Human Memory. Part 1: D’you remember?. Well, it depends…. The Big Three Questions. How does information get into memory? How is information maintained in memory? How is information pulled back out of memory?. Encoding, Storage, Retrieval. Encoding: Getting Information Into Memory.

592 views • 30 slides

Human Memory

Human Memory. Modules 26-30 (Myers). Human Memory: Basic Questions. How does information get into memory? How is information maintained in memory? How is information pulled back out of memory?. The Analogy. Figure 7.2 Three key processes in memory.

828 views • 56 slides

Human Memory

Human Memory . Encoding:Getting Information Into Memory. Three key processes of memory are encoding, storage, and retrieval. Encoding involves forming a memory code. Storage is maintaining encoded information in memory over time. Retrieval involves recovering information from memory.

366 views • 14 slides

Human Memory

Human Memory. It is good to have an end to journey towards; but it is the journey that matters, in the end. Ursula K. Le Gui. Memory. Process by which information is: Acquired Encoding Stored in the brain Storage Later retrieved Retrieval Eventually (possibly) forgotten.

514 views • 32 slides

Human Memory

Human Memory. Do you recognize a real penny?. Did you have to look?. Did you need to check a real penny to be sure? Nickerson and Adams (1979) conducted an experiment and concluded that most people are not able to pick the correct penny from this drawing (Weiten 1995).

1.78k views • 25 slides

Human Memory

Human Memory. What we usually think of as “memory” in day-to-day usage is actually long-term memory, but there are also important short-term and sensory memory processes, which must be worked through before a long-term memory can be established.

938 views • 8 slides

[Human Memory] 10.Knowledge

[Human Memory] 10.Knowledge

[Human Memory] 10.Knowledge. LEE JI HOON 2008.10.8 . Question. 1. How do we access the information in generic memory? 2. How to add up your knowledge?. What is knowledge?. Knowledge is what you know. Generic memory Everyday, ordinary knowledge

421 views • 18 slides

Human Memory

Human Memory. By Chris Gardner 10155520. The Different Types of Memory. Short Term Memory. Types of Long Term Memory. Types of Long Term Memory. Procedural Long Term Memory is, essentially, remembering skills like playing a guitar or riding a bike.

705 views • 14 slides

Human Memory

Human Memory . A basic overview of Human Memory, its primary functions and importance. Written by Zoe Lent 10138330. Fig A (University of California, 2009). Human Memory. What is memory?

452 views • 15 slides

Human Learning & Memory

Human Learning & Memory

Human Learning & Memory. Siena Heights University Chapters 7 & 8 Dr. S.Talbot. Cognitive Process Def. – how people ______, ______, ______ and ______ about their environment. Edward Tolman’s Purposive Behaviorism Learning can be internal rather than external (latent).

252 views • 13 slides

Human Learning & Memory

Human Learning & Memory. Siena Heights University Chapters 1 & 2 Dr. S.Talbot. Why is learning important? Definition of learning:a ______-term change in mental ________ or representations as a result of ________. The role of behavior in learning. New behaviors Frequency of behaviors

566 views • 26 slides

PSY 368 Human Memory

PSY 368 Human Memory

PSY 368 Human Memory. Short Term Memory. Announcements. Quiz 2 today Exam 1 Feb. 15 (Wed) Experiment 1 Report. Due Feb. 15 (so start it early, so that you can study for the exam). Experiment 1 assignment.

632 views • 54 slides

Chapter 7: Human Memory

Chapter 7: Human Memory. Human Memory: Basic Questions. Many psychologists study factors that help or hinder memory storage and retrieval…thus attempting to answer 3 basic questions… How does information get into memory? (ENCODING) How is information maintained in memory? (STORAGE)

872 views • 55 slides

PSY 368 Human Memory

PSY 368 Human Memory. Memory Processing. Announcements. Craik and Lockhart (1972) download, read, and answer focus questions, due Monday Feb 27 Processing views homework due Web Feb 29. Structural Model. Memory composed of storage structures that hold memories for a period of time

807 views • 34 slides

PSY 368 Human Memory

PSY 368 Human Memory. Reconstructive Memory cont. Announcements. Experiment 3 Report due April 16 If you missed the details of the Experiment, I included them in the lectures last week. Experiment 3.

590 views • 45 slides

PSY 368 Human Memory

PSY 368 Human Memory. Semantic Memory cont. Reconstructive Memory. Announcements. Data from Experiment 3 due April 9 (Mon) Experiment 3 Report due April 16 If you missed the details of the Experiment, I included them again at the end of this lecture

1.64k views • 137 slides

Chapter 9-Human Memory

Chapter 9-Human Memory

Chapter 9-Human Memory. Memory Constructed. Flashbulb memory Old teaching: New understanding. Car Accident video. 3 Questions to begin this chapter:. How does information get into memory?. Encoding. How is information maintained in memory?. Storage.

836 views • 57 slides

PSY 368 Human Memory

PSY 368 Human Memory. Development of Memory. Development of Memory. Outline for this week Studying infants & children Basic Processes and Capacities Methodological issues Memory in the Elderly What abilities decline? Why do they decline?. Studying Young Kids. Iconic memory

579 views • 44 slides

PSY 368 Human Memory

PSY 368 Human Memory. Neuropsychology & Memory Review for Exam 2. Announcements. Focus Questions for Weldon and Roediger (1987) Due Monday Today Exam 2 Wednesday (March 28). Alzheimer ’ s Disease. Alzheimer ’ s disease Cortical, progressive dementia. Criteria

743 views • 52 slides

Chapter 7: Human Memory

Chapter 7: Human Memory. Human Memory: Basic Questions. How does information get into memory? How is information maintained in memory? How is information pulled back out of memory?. Figure 7.2 Three key processes in memory. Encoding: Getting Information Into Memory.

1.17k views • 34 slides

COMMENTS

  1. The human memory

    2. MEMORY INTRODUCTION Memory is the ability to encode, store and remember information and past experiences in the brain. Encoding: a process of making mental representation of information. It can also mean transferring from short term to long term. Storing: Process of placing encoded information into relatively permanent storage for later ...

  2. Human Memory (Psychology)

    Human MEMORY. 2. Let's define memory… •is an organism's ability to store, retain, and recall information and experiences. •is our ability to encode, store, retain and subsequently recall information and past experiences in the human brain. 3. 4. *Sensory Memory -is the shortest-term element of memory. -The ability to look at an item for ...

  3. PDF MEMORY

    b. Episodic memory is a long-term memory system that stores in-formation about specific events or episodes related to one's own life. 1. episodic memory is used to recall past events, such as a movie you saw last week, the dinner you ate last night, the name of the book your friend recommended, or a birthday party you attended.

  4. psychology of memory

    Models of Memory Formation The Atkinson-Shiffrin Model (1968) 1. Stimuli are recorded by our senses and held briefly in sensory memory. 2. Some of this information is processed into short-term memory and encoded through rehearsal . 3. Information then moves into long-term memory where it can be retrieved later.

  5. Memory Models in Psychology

    3. Baddeley's model of Working memory: With the glaringly obvious role of attention in manipulating information in working memory, Baddely created a model that better accounts for manipulation in working memory. There is an addition of 3 important features to the vague idea of short-term memory and working memory.

  6. The human memory—facts and information

    Memories. Humans retain different types of memories for different lengths of time. Short-term memories last seconds to hours, while long-term memories last for years. We also have a working memory ...

  7. Memory and the Brain

    Memory is not like a video recording. Human memory, even flashbulb memories, can be frail. Different parts of them, such as the time, visual elements, and smells, are stored in different places. When something is remembered, these components have to be put back together for the complete memory, which is known as memory reconstruction.

  8. PDF Human Memory

    Human memory: An introduction to research, data, and theory. 2d ed. Belmont, CA: Wadsworth. This textbook is aimed at advanced undergraduate students and provides a comprehensive foundation for the study of human memory, including an introduction to mathematical and computational models of human memory. Roediger, H. L., ed. 2008.

  9. What Is Memory?

    Memory refers to the psychological processes of acquiring, storing, retaining, and later retrieving information. There are three major processes involved in memory: encoding, storage, and retrieval. Human memory involves the ability to both preserve and recover information. However, this is not a flawless process.

  10. PPT

    Presentation Transcript. Human Memory • What we usually think of as "memory" in day-to-day usage is actually long-term memory, but there are also important short-term and sensory memory processes, which must be worked through before a long-term memory can be established. • The different types of memory each have their own particular ...

  11. Cognitive neuroscience perspective on memory: overview and summary

    This paper explores memory from a cognitive neuroscience perspective and examines associated neural mechanisms. It examines the different types of memory: working, declarative, and non-declarative, and the brain regions involved in each type. The paper highlights the role of different brain regions, such as the prefrontal cortex in working ...

  12. The Properties of Human Memory and Their Importance for ...

    Sensory memories are the memories which are stored for tiny time periods and which originate from our sensory organs (such as our eyes or our nose). They are typically retained for less than 500 milliseconds. Visual sensory memory is often known as iconic memory. Sensory visual memories are the raw information that the brain receives (via the ...

  13. Memory

    the memory is the power of Human being. what is memory and its orientation is given in this presentation. Memory. ... This ppt is about the human memory and also tell about the type of human memory. Also give information about level of processing. Memory - cognition .

  14. Human MEMORY.

    Presentation transcript: 1 Human MEMORY. 2 Let's define memory… is an organism's ability to store, retain, and recall information and experiences. is our ability to encode, store, retain and subsequently recall information and past experiences in the human brain. 4 -is the shortest-term element of memory. Sensory Memory -is the shortest ...

  15. 15 Presentations on the Human Memory

    The human memory and brain has been at the forefront of scientific research for hundreds of years, yet doctors, scientists and psychologists continue to discover more at an increasingly rapid pace. The speech by Sally Hogshead reveals to audience members that the average attention span of humans is only nine seconds long today (the same as fish ...

  16. PPT

    Presentation Transcript. Types of Long Term Memory • Procedural Long Term Memory is, essentially, remembering skills like playing a guitar or riding a bike. • Declarative Long Term Memory has two subsets, Semantic and Episodic. • Semantic memory is related towards words and word meaning, like remembering names or having a large vocabulary.

  17. Memory

    Study of the genetics of human memory is in its infancy though many genes have been investigated for their association to memory in humans and non-human animals. A notable initial success was the association of APOE with memory dysfunction in Alzheimer's disease. The search for genes associated with normally varying memory continues.

  18. Schedules of presentation and temporal distinctiveness in human memory

    Recency, in remembering a series of events, reflects the simple fact that memory is vivid for what has just happened but deteriorates over time. Theories based on distinctiveness, an alternative to the multistore model, assert that the last few events in a series are well remembered because their times of occurrence are more highly distinctive than those of earlier times. Three experiments ...

  19. Human Memory: Structures and Processes (2d ed.)

    cesses within the structures of human memory. The four chapters on long term memory emphasize semantic memory, encoding, retrieval, and for-getting. In chapter 8 the usual lucid presentation gives way to a lengthy discussion of models of semantic memory and theories of visual represen-tations.

  20. Human Memory

    Mar 28, 2011 • Download as PPT, PDF •. 78 likes • 18,697 views. Ryan Braganza. A Project on Human Memory for Psychology. Health & Medicine. 1 of 27. Download now. Human Memory - Psychology - Download as a PDF or view online for free.

  21. PPT PowerPoint Presentation

    PowerPoint Presentation. Human Memory. Don't ask too much of mere mortals. Material mainly from Dix et al chapter 1. * * * * * * * * * * * * * * Learning outcomes Describe the major categories of human memory Describe the major organization structures of long term memory How are these organization structures reflected in UI design * Memory ...

  22. The Human: Memory

    The Human: Memory - Download as a PDF or view online for free. The Human: Memory - Download as a PDF or view online for free. Submit Search. Upload. The Human: Memory ... Presentation on memory. Presentation on memory Jamil Ahmed AKASH ...

  23. Microsoft Copilot Studio: Building copilots with agent capabilities

    When developing copilots with agent capabilities, establishing clear boundaries is paramount. Copilots operate strictly within the confines of the maker-defined instructions, knowledge, and actions. The data sources linked to the copilot adhere to stringent security measures and controls, managed through the unified admin center of Copilot Studio.

  24. PPT

    Human Memory. Human Memory. Don't ask too much of mere mortals Material mainly from Dix et al chapter 1. Learning outcomes. Describe the major categories of human memory Describe the major organization structures of long term memory How are these organization structures reflected in UI design. Attention. 675 views • 21 slides