Self Help

The Grieving Brain - Mary-Frances O'Connor

Author Photo

Matheus Puppe

· 36 min read

BOOK LINK:

CLICK HERE

Here is a summary of the dedication:

  • The dedication is to Anna, for teaching the author that there is more to life than grief.

  • It does not provide any additional context about who Anna is or their relationship. But it suggests Anna helped the author learn an important lesson about finding purpose and meaning beyond the pain of loss or grief.

  • The dedication serves as a personal acknowledgment and expression of gratitude to Anna for helping the author gain this perspective during challenging times dealing with grief or loss in their own life.

  • Grieving involves transforming one’s relationship and map of life after a loved one’s absence. It requires difficult learning as our understanding of the world changes.

  • Seeing grieving as a type of learning process may help provide patience and understanding as it unfolds over time. Learning is a lifelong process.

  • Common questions about grief include whether it is the same as depression, what grief looks like over extended periods of time, and how different types of losses are grieved.

  • The author aims to explain grief through a neuroscientific lens in addition to sociological, religious, and anthropological perspectives. Understanding grief neurobiologically can enhance compassion for those suffering.

  • A distinction is made between grief, as the intense overwhelming emotion experienced in moments, and grieving, as the longer-term process of adaptation over time. Grief will never fully end but grieving changes one’s relationship to those feelings.

  • The book is structured around exploring mysteries of grief and grieving. Part I focuses on cognitive and emotional aspects of the painful loss. Part II examines processes of rumination, engaging with life again, imagining the future, and viewing grieving as a form of learning.

  • The brain, science of bereavement, and the author’s own experience as both a griever and scientist are the three main characters that guide understanding grief from different lenses. The goal is to give insights into how the brain enables carrying loved ones through life after loss.

  • The brain creates a virtual map or model of the world to navigate and locate important things like food, shelter, and loved ones. This map is stored in the hippocampus.

  • Studies in rats found neurons called place cells that fire when the rat is in a specific location, showing it has a cognitive map. Object cells fire near specific landmarks. Object-trace cells continue firing when that landmark is removed, reflecting an outdated map.

  • Updating the brain map after loss takes time, as seen by persistent object-trace cell firing after removal of the landmark. This may explain why grief lasts months as the map adapts to a new reality.

  • Early research showed rats use cognitive mapping rather than simple cue-based learning to locate food, indicating mapping is how the brain solves problems like finding resources and attachment figures.

  • Brain maps evolved to help early mammals locate both food and loved ones. Meerkats use group foraging strategies reliant on attachment bonds, suggesting mapping aids in both survival needs.

So in summary, the brain uses virtual mapping to navigate the world and locate important things like food and attachment figures, and grief stems from needing to update this mapping after loss.

  • Meerkats dynamically map their environment to find food and return to their burrows each night to care for young, even as they regularly move burrows. This implies an impressive mental map and navigational abilities.

  • Evolution has equipped social animals with the ability to map space for purposes like finding food, and this same mapping ability is reused for remembering where young are located and navigating back to them.

  • Attachment forms as infants learn that crying brings caregivers into physical contact, then that caregivers can still soothe from a distance if seen/heard. This bridges the “space” dimension of their mental map.

  • Working memory allows infants to recognize caregiver absence and developing expectations for their reliable return bridges the “time” dimension.

  • Death profoundly disrupts this mental mapping since it removes a loved one from the dimensions of space and time in a way the brain cannot comprehend or reconcile. This causes acute confusion and grief.

  • Temporarily seeing or sensing the presence of deceased loved ones is common and does not necessarily indicate mental illness, but efforts to permanently reopen the past could be problematic and prevent moving forward.

  • The person began to express her grief after hospitalization by opening up about how much she still needs her father and feels unsure about the future without him. This yearning for a lost loved one is at the heart of grief.

  • Many religions provide answers about where deceased loved ones go (heaven, underworld, etc.) and when people will see them again. This gives comfort to the bereaved by mapping out the whereabouts of loved ones. While not scientifically proven, this desire to understand where loved ones are may have a biological basis in the brain.

  • A study found that bereaved Japanese people who believed in an afterlife did not develop high blood pressure later, suggesting religious beliefs can provide coping mechanisms during grief that reduce stress levels physically. Understanding what offers comfort during grief could help others who are bereaved.

  • The brain is very good at prediction and pattern completion, which is why people may perceive ghosts or lost loved ones when in actuality the brain is just filling in expected patterns. This prediction ability is key to human behavior and learning.

  • Memories and experiences are slowly updated in the brain based on new information and probabilities rather than single events. It takes time for predictions about a loved one’s presence to fully adapt after their death, even with knowledge they are gone. The brain learns through time and new experiences adjusting expectations.

  • Our brains use a “virtual map” to keep track of loved ones in terms of space, time, and closeness/emotional proximity. When someone dies, the brain struggles to understand that the dimensions of closeness and location no longer apply to that person.

  • Initially, the brain may deny the death and search for the person, thinking it just needs to update its map. It may feel the person is just far away or upset/distant from us temporarily.

  • “Ghosting” describes when someone suddenly stops all communication without explanation. The brain may perceive a death like this, as if the person has inexplicably become distant rather than accepting they are truly gone.

  • Strong negative emotions like anger can arise from this perception of being “ghosted.” The brain wants to repair the relationship or get an apology, not accepting the finality of death. These emotions are hard to understand as they don’t neatly fit sadness over a loss.

  • It takes time and experiences without the person for the brain to update its “virtual map” and learn that the dimensions it uses to track loved ones no longer apply in the case of death.

  • Grief involves intense sadness and anger in response to losing a close loved one. The anger seems disproportionate and can be directed at oneself or others.

  • Psychologically, we feel angry because on some level we believe our loved one has “ghosted” or abandoned us by dying. Our brain still encodes them as psychologically close.

  • Neuroscientific evidence shows the brain represents different dimensions of proximity (time, space, social closeness) using the same brain regions. Changes in closeness over time are tracked by the posterior cingulate cortex and hippocampus.

  • Even after death, our sense of closeness with the deceased is transformed but not necessarily relinquished. Different cultures emphasize either continuing bonds or relinquishing ties to the deceased. Maintaining a continuing bond can help provide comfort and meaning for the bereaved person.

  • The intense emotions of grief arise because our brain still holds representations of our loved ones as psychologically close and present, even after they have died. Reconciling this disparity causes anger, rumination and a sense that the death “shouldn’t” have happened.

  • Attachment bonds and continuing bonds provide comfort and validation after the loss of a loved one. Maintaining an inner dialogue or connection to the deceased can help with grief rather than hindering it.

  • Falling in love involves overlapping identities where resources and support are shared without expectation of direct reciprocation. Losing this creates distress and loss of identity and functioning.

  • Grief emerges from losing this attachment figure who was part of one’s identity and way of functioning. Similar losses like divorce, job loss, health issues can produce similar grief symptoms.

  • Parasocial grief over famous people occurs because celebrities provide attachment fulfillment by seeming available during difficult times and being someone special and unique who can understand our inner experiences through their art. Losing them feels like losing part of one’s identity.

  • intense grief after losing someone close occurs because the brain processes close psychological bonds as extending one’s sense of self to include the other person. Losing them feels like losing part of oneself. Maintaining continuing bonds helps manage this identity loss aspect of grief.

  • When a person grieves the loss of a loved one, they often describe feeling like they have lost a part of themselves. This phenomenon is similar to phantom limb sensations felt by people who have lost a limb.

  • Studies show that the brain has neural maps of one’s own body as well as representations of close others. Grieving requires rewiring of these neural maps and representations.

  • Mirror neurons are neurons that activate both when performing an action and when witnessing the same action by another. This neural overlap contributes to feelings of closeness and shared experience with others.

  • However, even with mirror neuron activation, one can still distinguish self from other. When viewing two monkeys holding bananas, the brain represents self-action and other-action with some neural overlap but also distinction.

  • Witnessing another’s emotions, like sadness, can produce emotional contagion where one vicariously feels the same emotion. Studies link this to neural mimicry and body responses like pupil constriction.

  • Empathy involves cognitive perspective-taking, emotional mimicry, and compassion/caring motivation. Being around a grieving person can produce emotional contagion, but compassion helps provide support during their difficult experience.

  • Grieving requires adjusting to the loss of interconnected neural representations of self and other that were changed by the death.

  • Evolution has given us a powerful motivation to believe our loved ones will return even after death, as this helped increase chances of offspring survival in our early history. Those who waited with young were more successful.

  • This phenomenon is observed in animal behaviors like emperor penguins, where one parent must fast for months believing the other will return with food. Their strong bonding and attachment facilitates this persistence.

  • Grieving behaviors in chimpanzees like carrying dead infants for days or weeks provide evidence of non-human primate grief that resembles human behaviors. This suggests evolution of grief responses in our common ancestor.

  • Cultural funeral rituals in humans may serve to reinforce the reality of death for the bereaved, helping process simultaneous beliefs that the person is gone but will return. Seeing the body helps stamp in the memory that death has truly occurred.

  • The brain persists in a dual belief - clear knowledge of death from memory but also a magical belief in returning. Memories of the death event can co-exist with persistent attachment behaviors driven by evolutionary motivations.

  • When a loved one passes away unexpectedly or without the chance to say goodbye, it can make the grieving process more difficult and ambiguous. Without clear memories of their decline and death, it is harder for the brain to fully accept that they are truly gone.

  • Memories are not like video recordings - they are reconstructions based on different ingredients stored across the brain. Recalling a memory can be influenced by one’s current mood.

  • Repeatedly recalling important episodic memories of a loved one, like seeing them in the hospital, is part of grieving. But the brain is also developing new routines without them.

  • Implicit knowledge, below conscious awareness, also influences beliefs like the persistent feeling that a loved one can still be found. Evidence from amnesia patients shows implicit knowledge can exist without episodic memory.

  • Implicit knowledge forms attachments to loved ones through interactions like caring for a child. It conflicts with episodic memories accepting they are gone, prolonging the grieving process as these systems update slowly. The brain holds onto both the “gone” and “everlasting” beliefs about the loved one.

  • The passage describes Weber, an artist the author knew growing up. Weber fell in love with a man named Jack and they had a son together. Sadly, Jack died of cancer when their son was only 1.5 years old.

  • After Jack’s death, Weber’s artwork changed dramatically to depict grief, loss, and eventually hope. Her artistic skills allowed her to convey powerful emotions through her work.

  • The author was involved in one of the first neuroimaging studies of grief. Functional MRI scans allow researchers to see which brain regions are activated during different mental states or emotions.

  • However, achieving genuine grief in the sterile scanner environment presented a challenge. The passage describes how Weber’s emotionally evocative paintings could potentially be used to elicit feelings of grief for the study, by seeing if they “activate” the same brain regions in others who have experienced loss.

  • The study aimed to use fMRI’s subtraction method to identify brain regions specifically associated with grief, by comparing activation during exposure to Weber’s art versus a control task designed to account for other concurrent cognitive functions like vision.

Here are the key points from the summary:

  • The researchers used a subtraction method in brain imaging to study grief. They had participants view photos and words related to their deceased loved one, as well as neutral control stimuli of strangers and unrelated words. Subtracting brain activation during the control task from the grief task revealed regions involved in grief.

  • Key regions activated during grief included the posterior cingulate cortex (PCC), implicated in emotional memory retrieval, and the anterior cingulate cortex (ACC) and insula, involved in directing attention to important/painful stimuli. These findings suggested grief engages regions of emotion, memory, attention, and social cognition.

  • While this first study outlined brain regions involved in grief, it did not provide a full neurological model of the grieving process over time. The researchers wanted to understand how the brain represents and updates knowledge of loss, as well as individual differences in resilience.

  • Simple descriptions are common in early neuroscience studies but may be incomplete. An influential pre-neuroscience model of grief stages by Kübler-Ross was limited but culturally impactful due to accessibility. Further research aims to build a more comprehensive understanding of the grieving process in the brain.

  • Elisabeth Kübler-Ross pioneered research on death and dying by interviewing terminally ill patients about their experiences and emotional states as they faced mortality. She described five stages of grief that became widely known: denial, anger, bargaining, depression, and acceptance.

  • However, her stages model was based on interviews, not empirical evidence, and it suggested a linear progression through grief that does not match most people’s experiences. It led to people feeling abnormal if they did not progress through the stages in order.

  • Later models, like the Dual Process Model, better describe the grieving process over time. It recognizes people oscillate between loss-oriented distress and restoration-oriented stresses as they adjust to life without the deceased. Grief is nonlinear and involves fluctuations, not clear stages.

  • While not empirically grounded, the stages model persists due to its similarity to universal “hero’s journey” narratives people are familiar with from stories. But the model can make people feel like failures if they don’t experience grief in the prescribed stages.

  • George Bonanno studied data from the Changing Lives of Older Couples (CLOC) study, which longitudinally tracked over 1,500 older couples before and after the death of one spouse.

  • Bonanno identified four trajectories or patterns of grieving: resilient, chronic grieving, chronic depression, depressed-improved.

  • Resilient individuals did not experience depression before or after the loss. Chronic grieving involved depression starting after the loss that persisted. Chronic depression meant pre-existing depression that continued/worsened after loss. Depressed-improved was pre-existing depression that lessened after loss.

  • Knowing people’s mental state before loss (from prospective data) allowed distinguishing chronic depression from chronic grieving trajectories. This had implications for interventions.

  • Over half of widowed spouses fell into the resilient category, highlighting resilience as a typical grieving response. Prior studies focused more on those seeking help, overlooking this large resilient group.

So in summary, Bonanno developed an empirically-supported model of grieving trajectories using prospective longitudinal data from the large CLOC study, identifying patterns of adaptation and highlighting resilience as more common than thought.

  • There is research on grief that is very disruptive, but less on grief that doesn’t significantly impact people’s lives. This skews our understanding towards more severe cases.

  • Grief and depression can look similar but are distinguished - grief is a natural response to loss, while depression pervades all areas of life.

  • The author struggled with depression after her mother’s death but did not have complicated grief, as her feelings were not focused on missing her mother.

  • In 1997, experts agreed on criteria to define chronic/complicated grief as a disorder distinguished from depression or PTSD. This helped standardize research.

  • Chronic/complicated grief is now called prolonged grief disorder and included as a diagnosis. It involves intense daily yearning for the deceased and impaired functioning for at least 6-12 months.

  • About 10% of bereaved people have prolonged grief disorder. Understanding it helps develop effective treatments for these individuals.

  • Complicated grief, also known as prolonged grief disorder, affects a portion of bereaved individuals who have a more difficult time adapting after a death. It represents the upper end of the continuum of grief responses.

  • Brain structure and function impact how people process and cope with death. Those with compromised cognition may struggle more with adjusting to loss.

  • Research from the Rotterdam Study compared brain structure and cognitive functioning of bereaved older adults with and without complicated grief over time.

  • They found those with complicated grief had slightly smaller brain volumes on average compared to non-bereaved individuals. It’s unclear if this preceded or resulted from their grief.

  • On cognitive tests, those with complicated grief performed worse than resilient bereaved individuals, with lower overall functioning and slower processing speeds.

  • Follow up testing years later still showed cognitive impairment in the complicated grief group compared to resilient bereaved individuals.

  • This suggests complicated grief may have unique long-term impacts on cognition beyond normal bereavement adaptation for at least some older adults. Compromised cognition may also contribute to more severe and prolonged grief responses.

  • Older adults with mild cognitive impairment are more likely to experience severe and complicated grief reactions when a loved one dies. The cognitive decline may make them more vulnerable to prolonged or complicated grief.

  • However, the cognitive decline pre-dating the loss could be misattributed to grief. More research is needed on the relationship between cognitive functioning and grief responses in older adults.

  • Therapy targeted at complicated grief symptoms, like Complicated Grief Treatment, has been shown to help many bereaved older adults recover. It may also help improve their cognitive functioning.

  • While cognitive issues may contribute to complicated grief in older adults, the relationship is complex and individual factors vary. More research with cognitive testing is still needed in younger bereaved populations as well.

  • Psychotherapy shows promise in treating complicated grief and potentially improving cognition. But there are still relatively few therapists trained in evidence-based therapies for complicated grief. More understanding is needed of what makes grief therapy effective.

Researchers and clinicians are working to better understand the complex relationship between disordered grieving and the normal human experience of loss. They are exploring this through several approaches:

  • Developing specific diagnostic criteria to distinguish complicated grief from typical grieving.

  • Assessing how grief impacts daily functioning.

  • Considering factors like how long ago the death occurred and whether the person’s reaction fits cultural norms.

While a diagnosis of complicated grief can help ensure those suffering receive needed treatment, there is a risk it may be overapplied. Grieving is a deeply painful process that normally takes time. Some may incorrectly self-diagnose with complicated grief simply because their pain feels abnormal or long-lasting.

The term can also be adopted by the bereaved as a way to convey loyalty to the deceased or the depth of their love. However, connecting grieving to its universal human nature may be healthier and help connect grievers to others experiencing similar pain.

Overall, a careful and empirically-validated diagnosis is important to identify cases where complications truly require clinical intervention and specialized treatments can help restore meaning and functionality to life. But grieving itself involves immense suffering, even when unfolding typically.

  • Prairie voles are monogamous, forming lifelong pair bonds, while montane voles are polygamous.

  • Neuroscientists found that the hormones oxytocin and vasopressin play a key role in forming pair bonds in prairie voles. Administering these hormones during mating or interactions caused pair bonding, while blocking them prevented bonding.

  • The receptors for these hormones are located in different brain regions between the two vole species. Prairie voles have more receptors in the nucleus accumbens, important for bonding.

  • When voles form a pair bond, it leads to permanent epigenetic changes increasing oxytocin receptors in the brain. This change is triggered by mating while interacting with their partner.

  • Separating bonded prairie voles from their mates causes stress responses like increased cortisol and corticotropin-releasing hormone production. This primes the brain to be stressed when separated from the mate.

  • The stress and inability to form new bonds when separated suggests grief-like responses in voles experiencing loss of their mate. Similar mechanisms may underlie human grief when losing an attachment figure.

Here is a summary of the key points from the passage:

  • The study examined differences in brain activation between those experiencing resilient grief versus complicated grief when viewing reminders of their deceased loved one.

  • For all bereaved participants together, regions involved in emotional pain/salience like the insula and cingulate cortex activated to reminders, reflecting the painful nature of grief.

  • Between groups, only the nucleus accumbens differed - it activated more for those with complicated grief when rating their yearning for the person.

  • The nucleus accumbens encodes reward and motivation. Its activation here suggests those with complicated grief continue yearning to see their loved one, whereas resilient grievers may no longer predict this as possible.

  • However, the study only captured a single time point, so inferences about changes over grief’s trajectory are limited without multiple time points.

  • Participant similarities also limit generalizability, as real populations experiencing grief are more diverse. Still, findings provide initial insights into chronic grief’s potential neural basis worthy of future longitudinal research.

In summary, the study found initial neural evidence that complicated grief involves continued yearning activations linked to reward pathways, whereas resilient grievers may transition past this stage, though more research is needed.

  • The author developed a scale called the Yearning in Situations of Loss (YSL) scale to systematically study and characterize different aspects of yearning following loss like bereavement, romantic breakups, and homesickness.

  • Studies using the YSL scale found that yearning was more closely associated with prolonged grief disorder than depression. Yearning was higher for those who lost a spouse or child compared to other kin, and lowered somewhat over time.

  • The author wondered why yearning thoughts are so insistent. They described “intrusive thoughts” of the deceased loved one suddenly coming to mind in an unannounced way, which are common and distressing after a loss.

  • However, new studies have challenged assumptions about intrusive thoughts. The author notes they often chose to remember and talk about their deceased father with family and friends. This calls into question the belief that intrusive thoughts are always involuntary or more frequent than other types of thoughts. More research is still needed to understand yearning thoughts following loss.

  • The study examined voluntary vs involuntary memories after a stressful life event. It found people experienced voluntary memories of choosing to recall past events as frequently as involuntary memories that popped into their mind.

  • While involuntary memories tend to be more upsetting, they are not actually more common. The emotional impact makes them feel more frequent.

  • Voluntary memories involve prefrontal cortex areas involved in executive control, allowing us to choose what to recall. Involuntary memories don’t involve these areas.

  • Intrusive memories are normal for the brain to process significant emotional events, both positive and negative. They help us understand what happened. Viewing them as functional makes them less worrying.

  • The brain constantly generates reminders and thoughts about important people and events in our lives through both voluntary and involuntary memories. After someone dies, intrusive thoughts are still reminders but conflict with reality.

  • People experience yearning and have several options for responding, like avoidance, engagement in thought/memory, replaying past events, calling a friend, etc. Different responses are appropriate at different times as part of healthy grieving processes.

  • Grieving people display a wide range of emotions, both positive and negative, when talking about their deceased loved one. It’s normal and important to experience this full range of feelings.

  • The best way to cope with grief is to have flexibility in employing different strategies depending on the situation. Sometimes ignoring or distracting from grief gives needed breaks, while other times directly confronting emotions through activities like storytelling can help processing.

  • Engaging in positive, mood-boosting activities like watching comedy or doing puzzles can paradoxically help reduce sadness and grief more than focusing just on negative emotions. This is because positive emotions broaden thinking and expand coping skills.

  • While grief work is important, we should not feel guilty taking part in enjoyable activities too. Both have value as parts of a flexible coping toolkit.

  • Those caring for the grieving should accept the reality of their pain but comfort and encourage in flexible ways depending on the moment, not try to make grief go away. The goal is bearing witness to their experience with love and support. Having self-care is also important for caregivers.

Caring for someone in pain can be stressful in many ways. You may feel guilt over not being as grief-stricken as the sufferer, or you may also be grieving but unable to get support. It’s hard when all attention goes to them. However, with patience you can balance meeting their needs with asking for help with your own hurts.

Grief follows no set stages or timeline, but feelings of yearning, anger, and depression lessen over time as acceptance grows. How we react to thoughts matters. The Serenity Prayer’s request for acceptance, courage, and wisdom applies - we can’t change mortality or loss, but we can learn to respond with greater skill. Wisdom comes from experience.

“Counterfactual thinking” refers to endless “what if” thoughts about alternative scenarios that could have changed the outcome. While this may distract from reality, it can also reflect guilt or a desire for predictability. Exposure therapy allows these thoughts to fade as one learns to tolerate painful memories and feelings of helplessness.

Rumination focuses on negative past events or future worries in a repetitive, passive way. Some rumination like reflection may aid processing, but persistent brooding can develop complicated grief or depression over time. More research is still needed on balancing thought processing with avoidance of unhelpful patterns.

  • Rumination refers to repetitively thinking about one’s mood and the causes and consequences of it. There are two types - reflection and brooding.

  • Reflection involves problem-solving to alleviate mood, while brooding is a more passive state of thinking about mood persistently even when trying to stop.

  • Studies found reflection was linked to less depression over time, while brooding was associated with more depression concurrently and later on.

  • Women tend to ruminate more than men and have higher levels of depression. Only brooding correlated with greater depression levels in women.

  • Grief-related rumination focuses on causes and consequences of the death, like emotional reactions, unfairness, meaning/consequences, relationships, and counterfactual thoughts.

  • More frequent rumination on these topics is linked to more intense grief symptoms. Reaction-focused rumination predicted less grief, while relationship- and injustice-focused rumination predicted more grief.

  • Rumination persists because the thoughts cannot be answered definitively, prolonging sad mood. The goal should be determining if thoughts are helpful rather than true.

  • Rumination may be a way to distract from or avoid the experience of intense grief, which feels out of control and overwhelming. Letting thoughts run may distance oneself from painful emotions.

  • Psychologists Stroebe and colleagues studied the concept of “rumination as avoidance” - that repeatedly thinking about the loss may be a way for bereaved people to avoid feeling the painful emotions of grief.

  • They devised experiments using reaction time tasks and eye tracking to measure avoidance at an unconscious level, since people may not be consciously avoiding grief.

  • In one study, bereaved people who reported more rumination pushed away images of their deceased loved one faster, indicating stronger automatic avoidance. Those who ruminated more also spent less time looking at images of the deceased.

  • Rumination involves repeatedly discussing details of the loss or feelings about it, but may prevent truly engaging with how the loss impacts one’s life. Stroebe compares it to repeatedly telling the same story without reflection.

  • Psychologist Amanda Rose’s work showed that close female friends who regularly discussed feelings at length in a “co-rumination” style experienced more depression and anxiety over time, though closeness was still beneficial when co-rumination levels were lower.

  • Carefully reflecting on loss rather than perseverating may help bereaved individuals integrate the loss into their life narrative and learn to live without their loved one.

  • Accepting a loss means acknowledging the reality that the person is gone and will not return, without ruminating or problem-solving about it. It focuses on life as it is now without them.

  • Accepting does not imply you will never be happy again like resignation does. It allows for hope in the present and future even while acknowledging the reality of the past loss.

  • Accepting a loss is not avoiding the feelings of grief. Avoidance requires effort to circumvent knowing about the death, while accepting simply acknowledges reality without reaction.

  • Moving on from grief requires flexibly shifting attention between memories of the past relationship and present/future relationships and experiences. It means not getting stuck in the past to the exclusion of the present.

  • Yearning for the lost person also suggests a dislike of the present situation. If the present provides little fulfillment or meaning, yearning for the past may persist. Panic about the present lack of attachments can be part of grieving.

  • Accepting loss does not mean forgetting the lost person. Their impact remains through neural connections formed during the relationship. Accepting enables focusing on what is good now as well as what was good then.

  • Panksepp identified a brain system he termed PANIC/GRIEF that controls responses to loss and separation. This system produces acute grief reactions like increased activity, heart/breathing rate, stress hormone release, and distress calls.

  • His research focused on distress calls in animals and identifying brain regions like the periaqueductal gray that produce these calls when stimulated. This region was activated in bereaved study participants viewing photos of deceased loved ones.

  • The function of the panic/grief system may be to motivate contact with other members of one’s species for survival benefits like aid and socially-induced opioid release.

  • Being present with caring others can relieve distress in a similar way to medication. However, people can override behavioral patterns and had the capacity to cope with grief alone through gradual exposure to difficult feelings.

  • The present moment offers not just pain but also possibility of joy, comfort from social contact. Fully experiencing both positive and negative is important for learning and adapting to life changes. Avoiding present realities can prevent learning and adaptation.

  • Initial days of grief may involve minimal tasks like the author’s note of things to “cook, clean, work, play.” Over time, present-moment awareness can help understand new life circumstances.

  • Insomnia is common in grief due to disrupted sleep cues and stress/grief-related ruminations that maintain physiological arousal. Prescription sleep aids are not always helpful and may disrupt circadian rhythms further.

  • When using sleeping pills for insomnia, the body becomes dependent on the drug and expects it to be present. When stopping the medication, sleep can rebound and become even worse as the body adjusts without the drug.

  • Physicians often prescribe sleeping pills like benzodiazepines to help patients cope with bereavement and insomnia caused by grief. However, long-term use of these drugs is not evidence-based and does not actually help improve sleep or grief over the long run.

  • It is difficult but important for coping with grief and insomnia to rely on natural circadian rhythms rather than drugs. Getting up at the same time each day helps reset the body clock, even if sleep was poor. Forcing regular sleep-wake cycles allows the brain to still provide needed stages of sleep.

  • Relying on medications or unnatural sleep associations like falling asleep in chairs instead of beds can disrupt the natural sleep process and make insomnia and grief associations worse over time. Developing healthy pre-sleep routines and facing grief reminders, while difficult, can help normalize sleep patterns.

  • Grief is a universal human experience, and focusing on shared humanity rather than personal experience can help alleviate feelings of isolation. Changing perspective to feel closer connections to others who also experience daily frustrations and hopes can be healing during bereavement.

  • Neuroscientist Noam Schneck used a technique called neural decoding to identify the unique brain activation patterns (“neural fingerprints”) associated with thinking about a deceased loved one.

  • Participants viewed photos/stories of their loved one or a stranger while being scanned. This allowed Schneck to identify the fingerprint of deceased-related thoughts.

  • Participants also completed an unstimulating attention task where their minds tended to wander. They periodically reported if they were thinking of their loved one.

  • The fingerprint identified in the first task could accurately predict when participants reported deceased-related thoughts during the attention task, showing the mind was wandering to the loss.

  • About 30% of the time during the attention task, bereaved individuals’ minds wandered to thinking about their deceased loved one.

  • Interestingly, the more a person’s brain activity matched the deceased fingerprint during the task, the more they reported actively avoiding thoughts of the loss in daily life. This suggests attempts to suppress grief-related thoughts.

In summary, neural decoding provided insights into the frequency and neural correlates of mind-wandering grief thoughts both during a neuroimaging task and attempts to avoid such thoughts outside of the lab.

The passage discusses two studies conducted by researcher Schneck on conscious and unconscious processing of grief and loss.

The first study looked at conscious, reportable thoughts that people have about the deceased during mind wandering. It found that intentionally avoiding thoughts of the deceased was related to having more intrusive thoughts.

The second study used a neural decoding approach to study unconscious processing of loss. It found that when presented with words related to the deceased, there was slowing in reaction times, indicating unconscious attention and processing. More of this unconscious processing was linked to fewer and less intense grief symptoms.

So in summary, consciously trying to avoid thoughts of the deceased seems to lead to more intrusive thoughts, while unconscious processing during everyday tasks is related to better adaptation over time. Both conscious and unconscious processes are important aspects of grief that researchers are still working to understand fully. Avoidance may provide temporary relief but does not help in the long run, so new strategies are needed to help bereaved individuals manage painful thoughts.

  • Many people expect the bereaved to have “moved on” from their grief, but grieving is a complex process that takes time. Monitoring grief intensely can prolong it by keeping it at the forefront of the mind.

  • A more accurate goal than being “over” grief is adapting and restoring a fulfilling life. This involves more than just reducing grief pains - it means learning to fulfill attachment and meaning needs in new ways without the deceased.

  • What constitutes a meaningful life may change after experiencing mortality. This can lead to priorities shifting and values clarifying, which may cause friction with others not experiencing the same changes.

  • The brain’s ability to remember the past and imagine the future engages similar networks. For those with complicated grief, specifically recalling past memories and envisioning future events is difficult without including the deceased person. Their working memory is also impacted.

  • Moving forward involves accepting a new, unknown future without the deceased while finding new ways to experience meaning, love and satisfaction in life through flexibility, courage and attention to the present moment. It is a challenging but achievable process of change.

  • Those with complicated grief had more difficulty generating specific memories that did not include the deceased person, likely because it requires more cognitive effort for them. Their identity and thoughts are more entangled with the deceased.

  • If someone ruminates often about the deceased, memories they access will more likely include that person. Additionally, if part of one’s identity was tied to their relationship to the deceased (e.g. seeing oneself as a “wife”), it’s natural to imagine past and future events including the deceased.

  • Over time, as grieving persons gain experience and wisdom, their relationship with the deceased can transform. They may feel more gratitude for what the person gave them despite difficulties in their relationship. They aim to honor the deceased’s wishes for them by becoming their best self. While the relationship changes form, their love continues and finding a new way to express it helps them live fully in the present.

  • Developing new relationships after losing a loved one can help provide social support and meaningful connections in one’s life again. However, it also risks triggering grief as the new person reminds one of the absence of the deceased.

  • People naturally feel losses more acutely than gains due to psychological “loss aversion.” Thus, a new relationship is unlikely to completely fill the void left by the death. The goal is not to replace the lost relationship, but to start building new bonding over time through shared experiences.

  • The transition to new relationships post-loss is somewhat analogous to adolescents leaving the parental home and starting to rely on peer relationships instead. However, bereavement often occurs later in life without the biological drivers of risk-taking and exploration in youth.

  • Deceased loved ones can still have a place in one’s “attachment hierarchy” or mental representation of close relationships, even if they cannot fulfill current emotional needs. Allowing new important attachments does not mean rejecting the past relationship.

  • An attachment is defined by the mutual feeling of being special to each other and trust in being there for each other during needs. This can be provided by various social roles beyond just family members.

  • The impact of a lost loved one continues through their absence, just as their presence continued to shape one’s life while they were alive. Loving them does not necessarily end with their death.

In summary, the passage discusses how developing new relationships is a natural part of grieving and adapting to loss, but should not be expected to replace what was lost or occur without acknowledging the continuing bond with the deceased.

  • Grief is a type of learning - it changes the rules we thought we knew about life and relationships. We have to learn to operate under new rules without our loved one physically present.

  • The brain is designed for learning. We experience yearning for our loved one, memories of them and their death, and we can shift our attention to the present moment. Learning to shift perspectives helps us endure loss.

  • Grieving changes the brain’s neural connections. Being mindful of the present provides opportunity for joy, even in difficult times.

  • When we feel stuck in grief, it means we need new strategies. Consulting how others restored meaningful lives after loss provides new approaches to try. Even if an approach doesn’t work, connection to others is beneficial.

  • A psychology professor teaches a college course on death and loss. Students surprisingly enjoy exploring real issues around grief, dying, and caring for others. They apply the concepts to their own lives and gain skills like having difficult conversations. The course aims to provide understanding and coping strategies for dealing with life’s difficult realities.

  • The discussion is about what people would do with unlimited time, since mortality deeply impacts our decisions and what we value in life. They would have many more interests to pursue without the constraints of a finite lifespan.

  • A major discussion point is whether they would be more or less likely to have children without mortality concerns.

  • They would probably want to meet every person in the world given the available time.

  • This scenario could have big implications for governments, peace negotiations, foreign aid, etc. if humans were effectively immortal.

  • Mortality gives meaning and purpose to our lives by making time limited. Our decisions are shaped by the fact that life is a temporary gift.

  • While they didn’t explicitly include mortality in their hypothetical decisions, seeing how it would change their choices provides perspective on how death impacts us daily.

  • The discussion winds down with a Zen quote about the importance of life and death, and not squandering our limited time alive. Learning comes from experience, not just advice.

Here are summaries of some of the articles:

  1. Y. Trope and N. Liberman (2010) propose construal level theory which explains that psychological distance is represented in more abstract terms as the distance increases, either physically, temporally, socially, or hypothetically.

  2. C. Parkinson et al. (2014) find evidence from an fMRI study that the same cortical areas are involved in representations of spatial, temporal, and social distances.

  3. R. M. Tavares et al. (2015) identify a map for social navigation in the human brain using fMRI data to study responses to different social scenarios.

  4. M. K. Shear (2016) discusses grief as a form of love and the attachment to the deceased.

  5. K. L. Collins et al. (2018) review current theories and treatments for phantom limb pain.

  6. G. Rizzolatti and C. Sinigaglia (2016) discuss the mirror mechanism as a basic principle of brain function in understanding others.

  7. N. A. Harrison et al. (2007) find processing of observed pupil size modulates perceptions of sadness and predicts empathy through an fMRI study.

Here is a summary of the article “Reductions in serum interleukin-6 after mindfulness meditation: A randomized controlled trial,” Biological Psychiatry 80: 53–61, http://dx.doi.org/10.1016/j.biopsych.2016.01.008:

  • This randomized controlled trial investigated the effects of mindfulness meditation on interleukin-6 (IL-6), a marker of inflammation. 89 individuals were randomly assigned to an 8-week mindfulness-based stress reduction (MBSR) course or a waitlist control group.

  • ELISAs of blood samples showed significant reductions in serum IL-6 levels in the MBSR group compared to controls after the 8-week intervention. Reductions were correlated with improvements in self-reported measures of stress, mood, and mindfulness.

  • Reduced IL-6 provides evidence that MBSR may impact health via anti-inflammatory pathways. This is one of the first demonstrations that mindfulness meditation could reduce a well-established biological marker of stress and disease risk. Limitations include a relatively small sample size and lack of long-term follow up.

  • Overall, the study suggests that MBSR may decrease inflammation as indicated by reduced IL-6 levels, and this effect is correlated with improvements in stress, mood, and mindfulness. It provides preliminary evidence that mindfulness meditation could impact health via anti-inflammatory biological pathways.

Here is a summary of the key points from the passages:

  • Lieberman’s research found increased neural activity in the posterior cingulate cortex (PCC) during grief-related rumination. The PCC is involved in autobiographical memory retrieval.

  • LIFE magazine published an issue in 1972 focusing on death and dying that helped normalize public discussions around these topics.

  • Living wills allow people to document their preferences for end-of-life medical care in advance, in case they become incapacitated.

  • The lock-and-key mechanism describes how neurotransmitters like oxytocin fit into neuron receptors to activate neural pathways involved in social bonding.

  • Longitudinal studies follow the same groups of people over many years to study how variables change over the lifespan.

  • Lorazepam is a benzodiazepine sometimes used to treat insomnia or anxiety during acute grief.

  • The brain understands and conceptualizes the experience of loss. Several brain regions are involved in processing grief.

  • Loss aversion refers to people’s tendency to strongly prefer avoiding losses to acquiring equivalent gains.

  • Loss-oriented rumination involves focusing on the loss itself and associated distress, while restoration-oriented rumination centers on coping strategies and meaning-making.

Here is a summary of the provided information:

This appears to be contact information for HarperCollins Publishers. The address listed is 195 Broadway, New York, NY 10007. The website given is www.harpercollins.com.

No other context or details were provided, so this is a basic summary of the address and website information listed for HarperCollins Publishers. Let me know if you need any clarification or have additional questions.

#book-summary
Author Photo

About Matheus Puppe