Author: Sooemrei

  • Syntax Meaning

    Verbal order declaring immoral levy.

  • Books of Samuel

    The Book of Samuel (Hebrew: ספר שמואל, romanized: Sefer Shmuel) is a book in the Hebrew Bible, found as two books (1–2 Samuel) in the Old Testament. The book is part of the Deuteronomistic history, a series of books (Joshua, Judges, Samuel, and Kings) that constitute a theological history of the Israelites and that aim to explain God’s law for Israel under the guidance of the prophets.

    According to Jewish tradition, the book was written by Samuel, with additions by the prophets Gad and Nathan, who together are three prophets who had appeared within 1 Chronicles during the account of David’s reign.Modern scholarly thinking posits that the entire Deuteronomistic history was composed b.c.630–540 BCE by combining a number of independent texts of various ages.

    The book begins with Samuel’s birth and Yahweh’s call to him as a boy. The story of the Ark of the Covenant follows. It tells of Israel’s oppression by the Philistines, which brought about Samuel’s anointing of Saul as Israel’s first king. But Saul proved unworthy, and God’s choice turned to David, who defeated Israel’s enemies, purchased the threshing floor where his son Solomon would build the First Temple, and brought the Ark of the Covenant to Jerusalem. Yahweh then promised David and his successors an everlasting dynasty.

    In the Septuagint, a basis of the Christian biblical canons, the text is divided into two books, now called the First and Second Book of Samuel.

    Biblical narrative

    The Jerusalem Bible divides the two Books of Samuel into five sections. Further subheadings are also based on subdivisions in that version:

    1 Samuel 1:1–7:17. Samuel
    1 Samuel 8:1–15:35. Samuel and Saul
    1 Samuel 16:1–2 Samuel 1:27. Saul and David
    2 Samuel 2:1–20:26. David
    2 Samuel 21:1–24:25. Supplementary Information

    1 Samuel

    Samuel (1:1–7:17)

    The childhood of Samuel (1:1–4:1a)

    A man named Elkanah, an Ephraimite from the city of Ramathaim-Zophim, has two wives, Peninnah and Hannah, the latter of whom is his favourite wife. A rivalry between the two develops based on the fact that Peninnah has children and Hannah does not. The childless Hannah vows to Yahweh, the lord of hosts, that if she has a son, he will be dedicated to God. Eli, the priest of Shiloh, where the Ark of the Covenant is located, thinks she is drunk, but when he realises she is praying, he blesses her. A child named Samuel is born, and Samuel is dedicated to the Lord as a Nazirite – the only one besides Samson to be identified in the Bible. Hannah sings a song of praise upon the fulfilment of her vow.

    Eli’s sons, Hophni and Phinehas, sin against God’s laws and the people, specifically by demanding raw rather than boiled meat for sacrifice and having sex with the Tabernacle’s serving women. However, Samuel grows up “in the presence of the Lord”: his family visits him each year, bringing him a new coat, and Hannah has five more children. Eli tries to persuade his sons to stop their wickedness but fails. As punishment for this, a holy man arrives, prophesying that Eli’s family will be cut off and none of his descendants will see old age.

    One night, God calls Samuel, and, thinking Eli is calling him three times, he rushes to Eli. Eli informs him that God wishes to speak to him, and God informs Samuel that the earlier prophecy about Eli’s family is correct. Samuel is initially afraid to inform Eli, but Eli tells him not to be, and that God will do what is good in His sight. Over time, Samuel grows up and is recognised as a prophet.

    The Ark in Philistine hands (4:1b–7:17)

    The Philistines, despite their initial worries when hearing the Israelite ritual of the entrance of the Ark of the Covenant, defeat the Israelites at the Battle of Aphek, capturing the Ark and killing Hophni and Phinehas, thus fulfilling the earlier prophecy. When Eli hears of these two events, particularly the capture of the Ark, he falls off his chair and dies. His daughter-in-law, in turn, goes into labour at this, and names her son Ichabod (‘without glory’) in commemoration of the capture of the Ark.

    Meanwhile, the Philistines take the Ark to the temple of their god Dagon, who recognizes the supremacy of Yahweh. The Philistines are afflicted with plagues, are unable to take the Ark into any city on account of the fear of the populations of those cities, and return the ark to the Israelites, but to the territory of the tribe of Benjamin, to the city of Beth Shemesh, rather than to Shiloh, from where it is passed to the city of Kiriath Jearim, where a new priest, Eleazar, son of Abinadab, is appointed to guard the ark for the twenty years it is there. The Philistines attack the Israelites gathered at Mizpah in Benjamin. Samuel appeals to God, the Philistines are decisively beaten, and the Israelites reclaim their lost territory. Samuel sets up the Eben-Ezer (the stone of help) in remembrance of the battle, and takes his place as judge of Israel.

    Samuel and Saul (8:1–15:35)

    The institution of the monarchy (8:1–12:25)

    In Samuel’s old age, he appoints his sons Joel and Abijah as judges but, because of their corruption, the people ask for a king to rule over them. God directs Samuel to grant the people their wish despite his concerns: God gives them Saul from the tribe of Benjamin, whom Samuel anoints during an attempt by Saul to locate his father’s lost donkeys. He then invites Saul to a feast, where he gives him the best piece of meat, and they talk through the night on the roof of Samuel’s house. Samuel tells Saul to return home, telling him the donkeys have been found and his father is now worrying about him, as well as describing a series of signs Saul will see on the way home. Saul begins to prophesy when he meets some prophets, confusing his neighbours. Eventually, Samuel publicly announces Saul as king, although not without controversy.

    Shortly after, Nahash of Ammon lays siege to Jabesh Gilead and demands that everyone in the city have their right eye gouged out as part of the peace treaty. The Jabeshites send out messengers, looking for a saviour. When Saul hears of the situation, he gathers a 330,000-strong army and launches a surprise attack at night, leading Israel to victory and saving Jabesh, thus proving those who doubted him wrong. Saul’s kingship is renewed.

    Samuel is aware he is the final judge and that the age of kings is about to begin, and speaks to the Israelites, demonstrating his innocence and recapping the history of Israel. He calls on the Lord to send thunder and rain, and rebukes the people for their desire for a king. Nonetheless, he tells them that as long as they refrain from idol worship, they will not perish – but if they do, calamity will befall the kingdom.

    The beginning of Saul’s reign (13:1–15:35)

    Despite his numerous military victories, Saul disobeys Yahweh’s instructions. First of all, after a battle against the Philistines, he does not wait for Samuel to arrive before he offers sacrifices. Meanwhile, it turns out that the Philistines have been killing and capturing blacksmiths in order to ensure the Israelites do not have weapons, and so the Israelites go to war essentially with sharpened farm instruments. Saul’s son Jonathan launches a secret attack by climbing a pass into the Philistine camp and kills twenty people in the process. The panic this creates leads to a victory for the Israelites. Jonathan finds some honey and eats it, despite a royal decree not to eat until evening.

    Jonathan begins to doubt his father, reasoning an even greater victory could have been achieved if the men had eaten. The royal decree has other unintended knock-on effects, namely that the men start killing and eating animals without draining the blood. To counteract this, Saul sets up an altar so the proper laws can be observed. When a priest suggests asking God before launching another attack, God is silent, leading Saul to set up a pseudo-legal procedure to ascertain whose fault it is that God has abandoned them. The lot falls on Jonathan, but the men refuse to let him be executed since he is the reason for their victory.

    Over time, Saul fights the Moabites, the Ammonites, the Edomites, the Zobahites, the Philistines and the Amalekites, winning victory over them all. His kingdom is in a constant state of war, and he constantly recruits new heroes to his army. However, he disobeys God’s instruction to destroy Amalek: Saul spares Agag, the Amalekite ruler, and the best portion of the Amalekite flocks to present them as sacrifices. Samuel rebukes Saul and tells him that God has now chosen another man to be king of Israel. Samuel then kills Agag himself.

    Saul and David (16:1–31:13)

    David at court (16:1–19:7)

    Samuel travels to Bethlehem to visit a man named Jesse, with God promising Samuel can anoint one of his sons as king. However, while inspecting Jesse’s sons, God tells Samuel that none of them are to be king. God tells Samuel to anoint David, the youngest brother, as king. Saul becomes ill and David comes to play the harp to him. Saul takes a liking to David and David enters Saul’s court as his armor-bearer and harpist.

    A new war against the Philistines begins, and a Philistine champion named Goliath emerges, challenging any Israelite to one-on-one combat, with the loser’s people becoming subject to the winner. David goes to take food to his brothers in the Israelite camp, learns of the situation and the reward Saul is willing to give to the person who kills him great wealth, his daughter’s hand in marriage and exemption from taxes for the killer’s family and tells Saul he will kill Goliath. Saul wants him to wear his armour, but David finds he cannot because he is not used to it. Seeing David’s youth, Goliath begins to curse him. David slings a stone into Goliath’s forehead, and Goliath dies. David cuts off Goliath’s head with Goliath’s sword.

    Jonathan befriends David. Saul begins to send David on military missions and quickly promotes him given his successes, but begins to become jealous of David after the Israelites make up a song about how much more successful David is than Saul. One day, Saul decides to kill David with a spear, but David avoids him. Saul realises that God is now with David and no longer with him, making him scared of David. He therefore seeks other ways to pacify David. First, he sends him on military campaigns, but this only makes him more successful.

    Next, he tries to marry him off to his daughter Merab, but David refuses, and so Merab is married off to the nobleman Adriel. However, Michal, another of Saul’s daughters, is in love with David. Although David is still unsure about becoming son-in-law to the king, Saul requires only 100 Philistine foreskins as dowry. Although this is a plan to have David captured by the Philistines, David kills 200 Philistines and brings their foreskins back to Saul.

    Saul then plots David’s death, but Jonathan talks him out of it.

    The flight of David (19:8–21:16)

    Once again Saul tries to kill David with his spear, and so David decides to escape, lowered out of a window by Michal, who then takes an idol, covers it in clothes and places goat’s hair on its head to cover David’s escape. David visits Samuel. When Saul finds this out, he sends men to capture David, but when they see Samuel they begin prophesying, as does Saul when he tries to capture David himself.

    David then visits Jonathan, and they argue about whether Saul actually wants to kill David. David proposes a test: he is to dine with the king the following day for the New Moon festival. However, he will hide in a field and Jonathan will tell Saul that David has returned to Bethlehem for a sacrifice. If the king accepts this, he is not trying to kill him, but if he becomes angry, he is. Jonathan devises a code to relay this information to David: he will come to the stone Ezel, shoot three arrows at it and tell a page to find them. If he tells the page the arrows are on his side of the stone, David can come to him, but if he tells them they are beyond the stone, he must run away. When Jonathan puts the plan into action, Saul attempts to kill him with his spear. Jonathan relays this to David using his code and the two weep as they are separated.

    David arrives at Nob, where he meets Ahimelech the priest, a great-grandson of Eli. Pretending he is on a mission from the king and is going to meet his men, he asks for supplies. He is given the showbread and Goliath’s sword. He then flees to Gath and seeks refuge at the court of King Achish, but feigns insanity since he is afraid of what the Philistines might do to him.

    David the outlaw (22:1–26:25)

    David travels to the cave of Adullam near his home, where his family visit him, until he finds refuge for them at the court of the king of Moab in Mizpah.

    One of Saul’s servants, Doeg the Edomite, saw David at Nob, and informs Saul that he was there. Saul arrives at the town, concludes that the priests are supporting David and has Doeg kill them all. One priest gets away: Abiathar, son of Ahimelech, who goes to join David. David accepts him, since he feels somewhat responsible for the massacre. David liberates the village of Keilah from the Philistines with the help of God and Abiathar. When God tells him that Saul is coming and the citizens of Keilah will hand him over to Saul, David and his men escape to the desert of Ziph, where Jonathan comes and recognises him as the next king. Some Ziphites inform Saul that David is in the desert, but Saul’s search is broken off by another Philistine invasion.

    After the invasion, Saul learns David is now living in the desert of En Gedi and resumes his search for him. At one point, he enters a cave to relieve himself. David and his men are further back in the cave. They discuss the possibility of killing Saul, but David opts to merely cut a corner off his robe and use this as proof that he does not in fact wish to kill Saul. Saul repents of how he has treated David, recognises him as the next king and makes him promise not to kill off his descendants.

    Samuel dies, and, after mourning him, David moves on to the Desert of Paran. Here he meets the shepherds of a Calebite named Nabal, and his men help protect them. At sheep-shearing time, he sends some of his men to ask for food. Nabal refuses, preferring to keep his food for his household. When his wife, Abigail, hears of this, she takes a large amount of supplies to David herself. This turns out to be at exactly the right moment, since David had just threatened to kill everyone in Nabal’s home. Abigail begs for mercy, and David agrees, praising her wisdom. That night Nabal has a feast, so Abigail waits until morning to tell him what she has done. He has a heart attack and dies ten days later. David marries Abigail and a woman from Jezreel named Ahinoam, but in the meantime Saul has married David’s first wife, Michal, off to a nobleman named Palti, son of Laish.

    Saul decides to return to pursuing David, and the Ziphites alert him as to David’s whereabouts. Saul returns to the desert of Ziph and sets up camp. One night, David and two companions, Achimelech the Hittite and Abishai son of Zeruiah (his nephew), go to Saul’s camp and find him asleep on the ground. Abishai advocates killing him, but David once again resists, content with taking a spear and water jug lying by Saul’s head. The next morning, David advises Abner, Saul’s captain, to put the soldiers to death for not protecting Saul, citing the absence of the spear and water jug as evidence. Saul interrupts, and once again repents of his hunt. He blesses David, David returns his spear and Saul returns home.

    David among the Philistines (27:1–31:13)

    David joins the Philistines out of fear of Saul, taking his wives with him and brutally destroying his enemies, largely the Geshurites, the Girzites and the Amalekites, but makes the Philistines believe he is attacking the Israelites, the Jerahmeelites and the Kenites instead. King Achish is pleased with him, and supposes he will continue to serve him. Eventually, the Philistines go to war with the Israelites, and David goes with them.

    Meanwhile, Saul is growing increasingly anxious about the upcoming battle, but cannot get advice from God. He decides to attempt to contact Samuel from beyond the grave. While he has expelled all the witches and spiritists, he learns that one remains at Endor. After Saul assures her she will not be punished, she agrees to summon Samuel. Samuel is not happy to be disturbed, and reveals that the Philistines will win the battle, with Saul and his sons dying in the process. Saul is shocked and, although at first reluctant, eats some food and leaves.

    Back in the Philistine camp, several of the rulers are not happy with the idea of fighting alongside David, suspecting he may defect during the battle. Achish therefore reluctantly sends David back instead of bringing him to Jezreel with the Philistine army. When David and his men arrive in Ziklag, they find it sacked by the Amalekites, and David’s wives taken captive. After seeking God’s advice, David decides to pursue the raiding Amalekites, finding the Egyptian slave of one, abandoned when he became ill, who can show them the band. When they are located and found to be feasting, David fights all day, with only 400 escaping on camels. David recovers everything and returns to the Besor Valley, where 200 men who were too exhausted to come with him have been guarding supplies. David announces all are to share in the treasure, and even sends some to the elders of Judah when he returns to Ziklag.

    Meanwhile, the Battle of Mount Gilboa is raging on and, as Samuel said, the Philistines are winning. Saul’s three sons have been killed, and he himself has been wounded by arrows. Saul asks his armor-bearer to run his sword through him rather than let him be captured by the Philistines, but does it himself when the armor-bearer refuses. When they see the battle going badly, the Israelites flee their towns, allowing the Philistines to occupy them. The next day, the Philistines find Saul, behead him, and take his armour to the temple of Astarte and his body to Beth Shan. When they hear what has happened, the citizens of Jabesh Gilead take his body and perform funerary rites in their city.

    2 Samuel

    Saul and David (continued) (1:1–1:27)

    David among the Philistines (continued) (1:1–1:27)

    Back in Ziklag, three days after Saul’s death, David receives news that Saul and his sons are dead. It transpires that the messenger is an Amalekite who, at Saul’s insistence, had killed Saul to speed his death along, and brought his crown to David. David orders his death for having killed God’s anointed. At this point, David offers a majestic eulogy, where he praises the bravery and magnificence of both his friend Jonathan and King Saul.

    David (2:1–20:26)

    David King of Judah (2:1–4:12)

    David returns to Hebron at God’s instruction. The elders of Judah anoint David as king, and as his first act he offers a reward to the people of Jabesh Gilead for performing Saul’s funerary rites. Meanwhile, in the north, Saul’s son Ish-bosheth, supported by Abner, has taken control of the northern tribes. David and Ish-bosheth’s armies meet at the Pool of Gibeon, and Abner and Joab, another son of Zeruiah and David’s general, agree to have soldiers fight in one-on-one combat. All this achieves is twelve men on each side killing each other, but a battle follows and David wins. During the Benjaminites’ retreat, Joab’s brother Asahel chases Abner and Abner kills him, shocking everyone. Joab and Abishai continue Asahel’s pursuit. A truce is declared when they reach a hill to avoid further bloodshed, and Abner and his men are able to cross the Jordan.

    The war continues as David builds a family. Meanwhile, the House of Saul is getting weaker. When Ish-bosheth accuses Abner of sleeping with Saul’s concubine Rizpah, Abner offers to join David, which David accepts as long as he brings Michal with him. At the same time, David sends a petition to Ish-bosheth for the return of Michal, which Ish-bosheth agrees to. Patiel follows her crying until he is told to return home. Following the return of Michal, Abner agrees to get the elders of Israel to agree to make David king. Joab believes Abner was lying in his purpose of coming to David and, after recalling him to Hebron, kills him in revenge for Asahel. David curses Joab’s family to always contain a leper, someone disabled or someone hungry. He then holds a funeral for Abner.

    By this point, the only other surviving member of Ish-bosheth’s family is Mephibosheth, Jonathan’s disabled son, who was dropped by his nurse as she attempted to escape the palace after the deaths of Saul and Jonathan. Ish-bosheth is murdered by Rechab and Baanah, two of his captains who hope for a reward from David, who stab him and cut off his head. They bring his head to David, but David has them killed for killing an innocent man. They are hanged by the pool of Hebron and Ish-bosheth’s head is buried in Abner’s tomb.

    David King of Judah and of Israel (5:1–8:18)

    David is anointed king of all Israel.

    Against all odds, David captures Jerusalem from the Jebusites. He takes over the fortress of Zion and builds up the area around it. Hiram I, king of Tyre sends craftsmen to build David a palace. Meanwhile, David’s family continues to grow. The Philistines decide to attack Israel now that David is king, but God allows David to defeat them in two battles, first in Baal Perizim and next in the Valley of Rephaim.

    The Ark is currently still in Baalah (another name for Kiriath Jearim), but David wants to bring it to Jerusalem. He puts it on a cart and employs the priests Uzzah and Ahio, both sons of Abinadab and brothers of Eleazar, to accompany it. A grand procession with musical instruments is organised, but comes to a sudden halt when the oxen stumble, causing Uzzah to touch the Ark and die. David is afraid to take it any further and stores it in the house of a man named Obed-Edom. When, after three months, Obed-Edom and his family have received nothing but blessings, David takes the Ark to Jerusalem. As part of the ceremony bringing the Ark into the city, David dances in front of it wearing nothing but an ephod. Michal sees this and is annoyed, but David says it was for the Lord, and thus it was not undignified. Michal never has any children.

    David wishes to build a temple, arguing that he should not be living in a palace while God lives in a tent. Nathan, a prophet, agrees. However, that night Nathan has a dream in which God informs him that David should not build him a temple for three reasons. Firstly, God has not commanded it, and has never complained about living in a tent before. Secondly, God is still working to build David and his house up and establish the Israelites in the Promised Land. Thirdly, God will establish one of David’s sons as king. He will build the temple, and his house will never be out of power. When Nathan reports this to David, David prays to God, thanking him for these revelations. David defeats the enemies of Israel, slaughtering Philistines, Moabites, Edomites, Syrians, and Arameans. He then appoints a cabinet.

    David’s family and the intrigues for the succession (9:1–20:26)

    Mephibosheth (9:1–9:13)

    David asks if anyone from the House of Saul is still alive so that he can show kindness to them in memory of Jonathan. Ziba, one of Saul’s servants, tells him about Mephibosheth. David informs Mephibosheth that he will live in his household and eat at his table, and Mephibosheth moves to Jerusalem.

    The Ammonite war and birth of Solomon (10:1–12:31)

    Nahash, king of Ammon dies and his son Hanun succeeds him. David sends condolences, but the Ammonites suspect his ambassadors are spies and humiliate them before sending them back to David. When they realise their mistake, they fear retaliation from David and amass an army from the surrounding tribes. When David hears that they are doing this, he sends Joab to lead his own army to their city gates, where the Ammonites are in battle formation. Joab decides to split the army in two: he will lead an elite force to attack the Aramean faction, while the rest of the army, led by Abisai, will focus on the Ammonites.

    If either enemy force turns out to be too strong, the other Israelite force will come to help their comrades. The Arameans flee from Joab, causing the Ammonites to also flee from Abishai. The Israelite army returns to Jerusalem. The Arameans regroup and cross the Euphrates, and this time David himself wins a decisive victory at Helam. The Arameans realise they cannot win, make peace with Israel and refuse to help the Ammonites again. The following spring, Joab destroys the Ammonites.

    While Joab is off at war, David remains in Jerusalem. One morning, he is standing on the roof of his palace when he sees a naked woman performing ablutions after her period. David learns her name is Bathsheba, and they have sex. She becomes pregnant. Seeking to hide his sin, David recalls her husband, Uriah the Hittite, from battle, David encourages him to go home and see his wife, but Uriah declines in case David might need him, and sleeps in the doorway to the palace that night. David, in spite of inviting Uriah to feasts, continues to be unable to persuade him to go home.

    David then deliberately sends Uriah on a suicide mission. David loses some of his best warriors in this mission, so Joab tells the messenger reporting back to tell David that Uriah is dead. David instructs Joab to continue the attack of the city. After Bathsheba has finished mourning Uriah, David marries her and she gives birth.

    Nathan comes to David and tells him a parable. In a town, there are a rich man and a poor man. The rich man has much livestock, but the poor man has only one lamb whom he loves like a child. One day, the rich man has a guest for dinner, and instead of slaughtering one of his own livestock, took the poor man’s lamb and cooked it. David angrily insists the rich man be put to death, but Nathan tells him he is the man, saying he has committed a sin to get something he already had plenty of (wives), and prophesies that his family will be gripped by violence, and someone will have affairs with his wives publicly.

    David repents, and Nathan tells him that while he is forgiven and will not die, his son with Bathsheba will. The child becomes ill, and David spends his time fasting and praying, but to no avail, because the child dies. David’s attendants are scared to tell him the news, worried about what he may do. He surprises everyone by ending his fasting, saying that he was fasting and praying was an attempt to persuade God to save his child, whereas fasting now isn’t going to bring the child back. After they have mourned, David and Bathsheba have another child, who they name Solomon (also called Jedediah).

    Back on the front line, in the city of Rabbah, Joab has gained control of the water supply. Joab invites David to finish capturing the city so that it may be named after himself. David gathers an army and travels up himself. He wins a victory, crowns himself king of the Ammonites, takes a large amount of plunder and puts the Ammonites into forced labour before returning to Jerusalem.

    Absalom (13:1–20:26)

    A complicated controversy begins to develop within the palace. Amnon, David’s son by Ahinoam, becomes lovesick for Tamar, David’s daughter by Maacah, daughter of Talmai, king of Geshur. Amnon’s advisor and cousin Jonadab suggests he pretend to be ill and ask Tamar to come and prepare bread for him so he can eat out of her hand. When she comes to his house, Amnon tells her to come to his bedroom. Here, after she refuses to have sex with him, Amnon rapes her. He then forces her to leave the house. She rips the gown which symbolises she is a virgin, puts ashes on her hand and walks around wailing. Tamar’s brother, Absalom, and David learn about this and become angry.

    Two years later, Absalom is shearing sheep at Baal Hazor and invites David and all his sons to come. David refuses, but blesses him and sends Amnon and the rest of his sons to him. Absalom holds a feast and gets Amnon drunk. He then instructs his servants to kill Amnon in revenge for his rape of Tamar. David’s other sons are disgusted and return to Jerusalem. David hears a rumour that Absalom has killed all of David’s sons, but Jonadab assures him that only Amnon is dead. Meanwhile, Absalom goes to live with his grandfather in Geshur for three years. After David has finished mourning Amnon, he considers visiting Absalom.

    Joab wants to help David, so he tells a wise woman from Tekoa to travel to Jerusalem pretending to be in mourning and speak to the king. The woman tells a story about her two sons, one of whom killed the other and whose death is now being called for. After some cajoling, David agrees to issue a decree ensuring that her son is not killed. The woman turns this back on David, and asks, then, why he has not forgiven his own son.

    After the woman admits that Joab put her up to this, David agrees to allow Absalom back to Jerusalem, but insists he does not come to the palace. Absalom becomes popular in Jerusalem due to his good looks. His family also grows during this time. Two years pass without Absalom being recalled to court. When Joab refuses to help him, Absalom sets his field on fire. This gets Joab’s attention, and finally Absalom manages to convince him to persuade David to allow him back to court.

    Absalom purchases a magnificent chariot, and begins campaigning to become a judge, principally by waiting outside the city gate, listening to the concerns of people coming to the king and pretending there is no-one to hear them, as well as embracing anyone who bows to him. Four years pass, and Absalom travels to Hebron, claiming to be fulfilling a vow, but in fact he hatches a plan to get the tribes of Israel to proclaim him king. The 200 guests who follow him do not know of his plan, and while he is at Hebron Absalom summons Ahitophel, David’s counselor.

    David is told of the increasing support for Absalom and decides to flee Jerusalem. He takes with him his wives and concubines, with the exception of ten, and a number of Cerethites, Pelethites and Gittites, led by a general named Ittai, who comes with David only after insisting on it. Abiathar and another priest named Zadok, together with a number of Levites who are guarding the Ark, also come, but go back when David tells them to return the Ark to Jerusalem. The procession climbs the Mount of Olives, where he meets his confidant Hushai the Arkite, who he sends back to Jerusalem to act as a spy, seeking to disrupt Ahitophel’s plans.

    On the other side of the mountain, David meets Ziba, who brings donkeys and fruit as supplies. He claims that Mephibosheth is hoping to be restored to the throne of Saul in the chaos, and David grants Ziba Mephibosheth’s estates. As the party approaches Bahurim, a Benjaminite named Shimei begins cursing and stoning David for the bloodshed he caused in the House of Saul. Abishai suggests executing him, but David considers that God has told Shibei to curse him and lets him carry on.

    Back in Jerusalem, Ahitophel and Hushai arrive at Absalom’s court. Absalom is at first suspicious of Hushai’s presence, but ultimately accepts him. Ahitophel suggests Absalom sleeps with David’s concubines who he left to take care of the palace in order to entrench the division between David and Absalom, so Absalom pitches a tent on the palace roof and does this in the view of all the Israelites. Ahitophel then suggests launching a sneak attack on David with 12,000 men. Hushai points out that David and his men are fighters, and that they could defeat the men, reducing morale. He suggests Absalom form a much larger army and lead it into battle himself.

    God has decided to frustrate Ahitophel’s advice so that Absalom can be defeated, so Absalom follows Hushai’s advice. Hushai then goes to Zadok and Abiathar and tells them to get word to David to cross the fords. Their sons, Ahimaaz and Jonathan, respectively, are staying at En Rogel, where they receive the message. Unfortunately, one of Absalom’s spies sees them so they have to hide in a well in Bahurim. The well’s owner’s wife hides them and lies to Absalom’s men that they have crossed the brook. After Absalom’s men are gone, the pair make it to King David and he manages to cross the Jordan in time.

    David and Absalom meet at Mahanaim, and David’s allies bring his army food, given his army is tired and exhausted after its time in the wilderness. David divides his army into thirds: one led by Joab, one led by Abishai and one led by Ittai. David intends to come out with his men, but his generals veto it. He decides to stay at the city, and instructs his generals to be gentle with Absalom. The battle is fought in the Wood of Ephraim. This proves to be a victory for David, in part because of the treacherous terrain. As Absalom meets David’s men, he passes under a tree. His long hair gets caught in the tree and he is hanged. Joab gets word of this, finds him and plunges three javelins into his heart, killing him. Joab declares the battle over and buries Absalom. Absalom’s monument is the pillar he built during his lifetime.

    Ahimaaz and a Cushite run to tell David the news of his victory and his son’s death. Ahimaaz declares the victory, but is not sure yet what the situation with Absalom is. The Cushite bears the same news, but also tells David that Absalom is dead. David begins to mourn, wishing he had died instead of Absalom. This prompts his men to start mourning as well, causing Joab to enter his tent in an attempt to talk sense into him. Joab points out that the battle has saved not only David’s life, but the lives of his wives and concubines, and thus it is humiliating for the men to have to mourn for the enemy. David agrees to come out and encourage the men.

    Given the sudden change in situation, the elders of Israel begin to argue about what to do next. David convinces the elders of Judah to escort him back to Jerusalem. They are joined by Shimei, who apologises to David. Abishai once again calls for the death penalty, but once again David grants clemency. Mephibosheth also comes to David, and explains the earlier situation: he had wanted to come with David and had told Ziba to saddle his donkeys, but Ziba had betrayed and slandered him. David offers to allow him and Ziba to split the land, but Mephibosheth allows Ziba to take the lot in celebration of David’s triumph.

    David invites his host in Mahanaim, Barzillai, to return to Jerusalem with him, but Barzillai protests on the basis that he is now eighty years old and thus will gain no enjoyment from coming. He gives David his servant Kimham in his place, and David promises to look after him. A scuffle breaks out between the Judahites and the other Israelites about why they specifically got to escort the king home. Attempting to resolve the issue, a Benjaminite named Sheba son of Bichri launches a rebellion against David, which all the tribes except Judah back.

    Back in Jerusalem, David begins to sort out the issues that were caused by his absence. First, he puts the ten concubines who were left behind into a guarded house and gives them pensions but does not sleep with them, allowing them to live the rest of their lives as widows. He then begins to sort out a defence against Sheba. He tells Amasa, the general whom he wishes to replace Joab, to summon the Judahite troops and have them in Jerusalem within three days, something he fails at. David therefore tells Abishai to start pursuing Sheba to effectively put down his rebellion before it has begun.

    Amasa meets Abishai and Joab at Gibeon. Amasa goes to meet Joab, but Joab’s dagger falls out of his tunic, stabbing Amasa in the stomach, killing him. He is covered with a cloth and placed in a field, and the army continues pursuing Sheba. They meet him at Abel Beth Maakah, a stronghold of Sheba’s rebellion, and begin to lay siege to it. A wise woman asks them why they want to destroy the city, and Joab responds they don’t want to destroy it, but merely end Sheba’s rebellion. The wise woman cuts off Sheba’s head and throws it to Joab from the city walls, thus ending the siege.

    Supplementary information (21:1–24:25)

    2 Samuel concludes with four chapters, chapters 21 to 24, that lie outside the chronological succession narrative of Saul and David, a narrative that will continue in The Book of Kings. Chapter 21 tells the story of a three-year long famine which takes place at the start of David’s reign. God explains this is a punishment for Saul’s genocide of the Gibeonites, a people group who are the remnants of the Amorites, whom Israel had promised to spare but Saul has massacred. David calls the Gibeonites and asks what he can do to make amends, hoping this will end the famine.

    The Gibeonites ask for seven of Saul’s descendants to kill, and David agrees. He spares Mephibosheth, but hands over Rizpah’s sons Armoni and Mephibosheth and the five sons of Merab and Adriel. They are killed by the Gibeonites and their bodies are exposed at the start of the barley harvest. Rizpah protects the bodies, and David agrees to take the bones of Saul, Jonathan and those killed by the Gibeonites and bury them in the tomb of Kish in Zelah. This pleases God and the famine ends.

    Another war then occurs with the Philistines. In the first battle, Abishai kills Ishbi-benob, a Philistine who had sworn to kill David, which leads to David’s army refusing to let him fight alongside them again for his own protection. The second battle takes place at Gob, and this time Sibbekai the Hushathite kills a Philistine named Saph. A third battle also takes place in Gob, where Elhanan, son of Jair kills Goliath’s brother. In the fourth battle, at Gath, Jonathan, son of Shimeah, kills a huge man with six fingers on each hand and six toes on each foot.

    Chapter 22 is similar to Psalm 18, and is a song David sang when he was delivered from Saul.

    Chapter 23 begins with David’s last words, a subdued speech in which David expresses gladness at the goodness of his house. It then tells stories of a group of men identified as ‘David’s Mighty Warriors’. Josheb-Basshebeth, Eleazar, son of Dodai and Shammah, son of Agee the Hararite all single-handedly win battles against the Philistines. One day, while David and his men are hiding in the cave of Adullam, David becomes homesick and, hearing the Philistines have taken over Bethlehem, cries out desiring water from Bethlehem’s well.

    These three men risk their lives to work their way through Philistine lines and bring water from the well back to David. David refuses to drink it and offers it to God because his warriors risked their lives for it. Abishai, we learn, achieved his high position by single-handedly killing three hundred men. Another warrior, Benaniah, son of Jehoiada, kills Moab’s two mightiest warriors, a lion, and a huge Egyptian with his own spear. The chapter finishes by listing David’s other mighty warriors, known as the Thirty.

    Chapter 24 tells the story of more calamities on Israel. God is angry once again at Israel, so he instructs David to take a census. Joab has his reservations, but ultimately relents. When the results come in, David realises what he has done, and begs God for mercy. Gad the prophet offers David three choices of punishment: three years of famine, three months of pursuit by his enemies or three days of plague. David chooses the plague. 70,000 people die.

    After three days the angel of death reaches Jerusalem, and is on the threshing floor of a man named Araunah the Jebusite, when God tells him to stop. David is horrified, arguing that it should be him and his family who are punished. Gad tells David to build an altar on the threshing floor of Araunah the Jebusite. Araunah offers to sell the land to David for free but David insists on paying. David pays fifty shekels of silver and builds the altar, stopping the plague.

    Composition

    Versions

    1 and 2 Samuel were originally (and, in most Jewish bibles, still are) a single book, but the first Greek translation, called the Septuagint and produced around the 2nd century BCE, divided it into two; this was adopted by the Latin translations used in the early Christian church of the West, and finally introduced into Jewish bibles around the early 16th century.

    In imitation of the Septuagint what is now commonly known as 1 Samuel and 2 Samuel, are called by the Vulgate, 1 Kings and 2 Kings respectively. What are now commonly known as 1 Kings and 2 Kings would be 3 Kings and 4 Kings in Bibles dating from before 1516. It was in 1517 that use of the division we know today, used by Protestant Bibles and adopted by Catholics, began. Traditional Catholic and Orthodox Bibles still preserve the Septuagint name; for example, the Douay–Rheims Bible.

    The Hebrew text that is used by Jews today, called the Masoretic Text, differs considerably from the Hebrew text that was the basis of the first Greek translation, and scholars are still working at finding the best solutions to the many problems this presents.

    Historical accuracy

    The Books of Samuel are considered to be based on both historical and legendary sources, primarily serving to fill the gap in Israelite history after the events described in Deuteronomy. According to Donald Redford, the Books of Samuel exhibit too many anachronisms to have been compiled in the 11th century BCE.

    Authorship and date of composition

    According to passages 14b and 15a of the Bava Basra tractate of the Talmud, the book was written by Samuel up until 1 Samuel 25, which notes the death of Samuel, and the remainder by the prophets Gad and Nathan. Critical scholars from the 19th century onward have rejected this idea. However, even prior to this, the medieval Jewish commentator Isaac Abarbanel noted that the presence of anachronistic expressions (such as “to this day” and “in the past”) indicated that there must have been a later editor such as Jeremiah or Ezra.Martin Noth in 1943 theorized that Samuel was composed by a single author as part of a history of Israel: the Deuteronomistic history (made up of Deuteronomy, Joshua, Judges, Samuel and Kings). Although Noth’s belief that the entire history was composed by a single individual has been largely abandoned, his theory in its broad outline has been adopted by most scholars.

    The Deuteronomistic view is that an early version of the history was composed in the time of king Hezekiah (8th century BCE); the bulk of the first edition dates from his grandson Josiah at the end of the 7th century BCE, with further sections added during the Babylonian exile (6th century BCE) and the work was substantially complete 550 BCE. Further editing was apparently done even after then. For example, A. Graeme Auld, Professor of Hebrew Bible at the University of Edinburgh, contends that the silver quarter-shekel which Saul’s servant offers to Samuel in 1 Samuel 9 “almost certainly fixes the date of this story in the Persian or Hellenistic period”.

    The 6th-century BCE authors and editors responsible for the bulk of the history drew on many earlier sources, including (but not limited to) an “ark narrative” (1 Samuel 4:1–7:1 and perhaps part of 2 Samuel 6), a “Saul cycle” (parts of 1 Samuel 9–11 and 13–14), the “history of David’s rise” (1 Samuel 16:14–2 Samuel 5:10), and the “succession narrative” (2 Samuel 9–20 and 1 Kings 1–2). The oldest of these, the “ark narrative,” may even predate the Davidic era.

    This view of late compilation for Samuel has faced serious scholarly opposition on the basis that evidence for the Deuteronimistic history is scant, and that Deuteronimistic advocates are not in consensus as to the origin and extent of the History. Secondly, the basic theological concerns identified with the Deuteronimistic school are tenets central to Hebrew theology in texts that are widely regarded as predating Josiah. Thirdly, there are notable differences in style and thematic emphasis between Deuteronomy and Samuel. Finally, there are widely acknowledged structural parallels between the Hittite suzerain treaty of the 2nd millennium BCE and the Book of Deuteronomy itself, far before the time of Josiah. The alternative view is that it is difficult to determine when the events of Samuel were recorded: “There are no particularly persuasive reasons to date the sources used by the compiler later than the early tenth century events themselves, and good reason to believe that contemporary records were kept (cf. 2 Sam. 20:24–25).

    Sources

    The sources used to construct 1 and 2 Samuel are believed to include the following:

    Call of Samuel or Youth of Samuel (1 Samuel 1–7): From Samuel’s birth his career as Judge and prophet over Israel. This source includes the Eli narrative and part of the ark narrative.
    Ark narrative (1 Samuel 4:1b–7:1 and 2 Samuel 6:1–20): the ark’s capture by the Philistines in the time of Eli and its transfer to Jerusalem by David – opinion is divided over whether this is actually an independent unit.
    Jerusalem source: a fairly brief source discussing David conquering Jerusalem from the Jebusites.
    Republican source: a source with an anti-monarchial bias. This source first describes Samuel as decisively ridding the people of the Philistines, and begrudgingly appointing an individual chosen by God to be king, namely Saul. David is described as someone renowned for his skill at playing the harp, and consequently summoned to Saul’s court to calm his moods. Saul’s son Jonathan becomes friends with David, which some commentators view as romantic, and later acts as his protector against Saul’s more violent intentions. At a later point, having been deserted by God on the eve of battle, Saul consults a medium at Endor, only to be condemned for doing so by Samuel’s ghost, and told he and his sons will be killed. David is heartbroken on discovering the death of Jonathan, tearing his clothes as a gesture of grief.
    Monarchial source: a source with a pro-monarchial bias and covering many of the same details as the republican source. This source begins with the divinely appointed birth of Samuel. It then describes Saul as leading a war against the Ammonites, being chosen by the people to be king, and leading them against the Philistines. David is described as a shepherd boy arriving at the battlefield to aid his brothers, and is overheard by Saul, leading to David challenging Goliath and defeating the Philistines. David’s warrior credentials lead to women falling in love with him, including Michal, Saul’s daughter, who later acts to protect David against Saul. David eventually gains two new wives as a result of threatening to raid a village, and Michal is redistributed to another husband. At a later point, David finds himself seeking sanctuary amongst the Philistine army and facing the Israelites as an enemy. David is incensed that anyone should have killed Saul, even as an act of mercy, since Saul was anointed by Samuel, and has the individual responsible, an Amalekite, killed.


    Court History of David or Succession narrative (2 Samuel 9–20 and 1 Kings 1–2): a “historical novel”, in Alberto Soggin’s phrase, telling the story of David’s reign from his affair with Bathsheba to his death. The theme is of retribution: David’s sin against Uriah the Hittite is punished by God through the destruction of his own family, and its purpose is to serve as an apology for the coronation of Bathsheba’s son Solomon instead of his older brother Adonijah. Some textual critics have posited that given the intimacy and precision of certain narrative details, the Court Historian may have been an eyewitness to some of the events he describes, or at the very least enjoyed access to the archives and battle reports of the royal house of David.


    Redactions: additions by the redactor to harmonize the sources together; many of the uncertain passages may be part of this editing.


    Various: several short sources, none of which have much connection to each other, and are fairly independent of the rest of the text. Many are poems or pure lists.

    Manuscript sources

    Four of the Dead Sea Scrolls feature parts of the books of Samuel: 1QSam, found in Qumran Cave 1, contains parts of 2 Samuel; and 4QSama, 4QSamb and 4QSamc, all found in Qumran Cave 4. Collectively they are known as The Samuel Scroll and date from the 2nd and 1st centuries BCE.

    The earliest complete surviving Hebrew copy of the books of Samuel is in the Aleppo Codex (10th century CE). The complete Greek text of Samuel is found in older manuscripts such as the 4th-century Codex Sinaiticus.

    Themes

    The Book of Samuel is a theological evaluation of kingship in general and of dynastic kingship and David in particular.The main themes of the book are introduced in the opening poem (the “Song of Hannah”): (1) the sovereignty of Yahweh, God of Israel; (2) the reversal of human fortunes; and (3) kingship. These themes are played out in the stories of the three main characters, Samuel, Saul and David.

    Samuel

    Samuel answers the description of the “prophet like Moses” predicted in Deuteronomy 18:15–22: like Moses, he has direct contact with Yahweh, acts as a judge, and is a perfect leader who never makes mistakes. Samuel’s successful defense of the Israelites against their enemies demonstrates that they have no need for a king (who will, moreover, introduce inequality), yet despite this the people demand a king. But the king they are given is Yahweh’s gift, and Samuel explains that kingship can be a blessing rather than a curse if they remain faithful to their God. On the other hand, total destruction of both king and people will result if they turn to wickedness.

    Saul

    Saul is the chosen one: tall, handsome and “goodly”, a king appointed by Yahweh, and anointed by Samuel, Yahweh’s prophet, and yet he is ultimately rejected. Saul has two faults which make him unfit for the office of king: carrying out a sacrifice in place of Samuel, and failing to exterminate the Amalekites, in accordance to God’s commands, and trying to compensate by claiming that he reserved the surviving Amalekite livestock for sacrifice.

    David

    One of the main units within Samuel is the “History of David’s Rise”, the purpose of which is to justify David as the legitimate successor to Saul. The narrative stresses that he gained the throne lawfully, always respecting “the Lord’s anointed” (i.e. Saul) and never taking any of his numerous chances to seize the throne by violence. As God’s chosen king over Israel, David is also the son of God (“I will be a father to him, and he shall be a son to me…” – 2 Samuel 7:14). God enters into an eternal covenant (treaty) with David and his line, promising divine protection of the dynasty and of Jerusalem through all time.

    2 Samuel 23 contains a prophetic statement described as the “last words of David” (verses 1–7) and details of the 37 “mighty men” who were David’s chief warriors (verses 8–39). The Jerusalem Bible states that last words were attributed to David in the style of Jacob and Moses. Its editors note that “the text has suffered considerably and reconstructions are conjectural”.

    1 Kings 2:1-9 contains David’s final words to Solomon, his son and successor as king.

  • Goliath

    Goliath is a Philistine giant in the Book of Samuel. Descriptions of Goliath’s immense stature vary among biblical sources, with various texts describing him as either 6 ft 9 in (2.06 m) or 9 ft 9 in (2.97 m) tall. According to the text, Goliath issued a challenge to the Israelites, daring them to send forth a champion to engage him in single combat; he was ultimately defeated by the young shepherd David, employing a sling and stone as a weapon. The narrative signified King Saul’s unfitness to rule, as Saul himself should have fought for the Kingdom of Israel.

    Some modern scholars believe that the original slayer of Goliath may have been Elhanan, son of Jair, who features in 2 Samuel 21:19, in which Elhanan kills Goliath the Gittite, and that the authors of the Deuteronomistic history changed the original text to credit the victory to the more famous figure David.

    The phrase “David and Goliath” has taken on a more popular meaning denoting an underdog situation, a contest wherein a smaller, weaker opponent faces a much bigger, stronger adversary.

    Biblical accounts

    In 1 Samuel 17, Saul and the Israelites are facing the Philistines in the Valley of Elah. Twice a day for 40 days, morning and evening, Goliath, the champion of the Philistines, comes out between the lines and challenges the Israelites to send out a champion of their own to decide the outcome in single combat, but Saul is afraid. David accepts the challenge. Saul reluctantly agrees and offers his armour, which David declines, taking only his staff, sling, and five stones from a brook.

    David and Goliath confront each other, Goliath with his armor and javelin, David with his staff and sling. “The Philistine cursed David by his gods”, but David replies:

    “This day the Lord will deliver you into my hand, and I will strike you down, and I will give the dead bodies of the host of the Philistines this day to the birds of the air and to the wild beasts of the earth; that all the earth may know that there is a god in Israel and that all this assembly may know that God saves not with sword and spear; for the battle is God’s, and he will give you into our hand.”

    —1 Samuel 17:46–47
    David hurls a stone from his sling and hits Goliath in the center of his forehead, Goliath falls on his face to the ground, and David cuts off his head. The Philistines flee and are pursued by the Israelites “as far as Gath and the gates of Ekron”. David puts the armor of Goliath in his own tent and takes the head to Jerusalem, and Saul sends Abner to bring the boy to him. The king asks whose son he is, and David answers:

    “I am the son of your servant Jesse the Bethlehemite.”

    —1 Samuel 17:58

    Composition of the Book of Samuel

    The Books of Samuel, together with the books of Joshua, Judges and Kings, make up a unified history of Israel which biblical scholars call the Deuteronomistic History. The first edition of the history was probably written at the court of Judah’s King Josiah (late 7th century BCE) and a revised second edition during the exile (6th century BCE), with further revisions in the post-exilic period. Traces of this can be seen in contradictions within the Goliath story, such as that between 1 Samuel 17:54, which says that David took Goliath’s head to Jerusalem, although according to 2 Samuel 5 Jerusalem at that time was still a Jebusite stronghold and was not captured until David became king.

    Structure of the David and Goliath narrative

    The Goliath story is made up of base-narrative with numerous additions made probably after the exile:

    Original story
    The Israelites and Philistines face each other; Goliath makes his challenge to single combat;
    David volunteers to fight Goliath;
    David selects five smooth stones from a creek-bed to be used in his sling;
    David’s courage strengthens others and eventually others defeat four other giants, possibly brothers, but relatives, reference 2 Samuel 21:15-22.
    David defeats Goliath, the Philistines flee the battlefield.
    Additions
    David is sent by his father to bring food to his brothers, hears the challenge, and expresses his desire to accept;
    Details of the account of the battle;
    Saul asks who David is, and he is introduced to the king through Abner.

    Textual considerations

    Goliath’s height

    The oldest manuscripts, namely the Dead Sea Scrolls text of Samuel from the late 1st century BCE, the 1st-century CE historian Josephus, and the major Septuagint manuscripts, all give Goliath’s height as “four cubits and a span” (6 feet 9 inches or 2.06 metres), whereas the Masoretic Text has “six cubits and a span” (9 feet 9 inches or 2.97 metres).Many scholars have suggested that the smaller number grew in the course of transmission (only a few have suggested the reverse, that an original larger number was reduced), possibly when a scribe’s eye was drawn to the number six in line 17:7.

    Goliath and Saul

    The underlying purpose of the story of Goliath is to show that Saul is not fit to be king (but that David is). Saul was chosen to lead the Israelites against their enemies, but when faced with Goliath, he refuses to do so; Saul is a head taller than anyone else in all Israel (1 Samuel 9:2), which implies he was over 6 feet (1.8 m) tall and the obvious challenger for Goliath, yet David is the one who eventually defeated him. Also, Saul’s armour and weaponry are apparently no better than Goliath’s:

    “David declares that when a lion or bear came and attacked his father’s sheep, he battled against it and killed it, [but Saul] has been cowering in fear instead of rising up and attacking the threat to his sheep (i.e., Israel).”

    David’s speech in 1 Samuel 17 can be interpreted as referring to both Saul and Goliath through its animal imagery. When this imagery is considered closely, David can be seen to function as the true king who manipulates wild beasts.

    Elhanan and Goliath

    In 2 Samuel 21, verse 19, the Hebrew Bible tells how Goliath’s brother was killed by “Elhanan the son of Jaare-oregim, the Bethlehemite”. The fourth-century BC 1 Chronicle 20:5 explains the second Goliath by saying that Elhanan “slew Lahmi the brother of Goliath”, constructing the name Lahmi from the last portion of the word “Bethlehemite” (“beit-ha’lahmi”), and the King James Bible adopted this into 2 Samuel 21:18–19, but the Hebrew text at Goliath’s name makes no mention of the word “brother”. Most scholars dismiss the later 1 Chronicles 20:5 material as “an obvious harmonization”.

    Goliath and the Greeks

    The armor described in 1 Samuel 17 appears typical of Greek armor of the sixth century BCE; narrative formulae such as the settlement of battle by single combat between champions has been thought characteristic of the Homeric epics (the Iliad) rather than of the ancient Near East. The designation of Goliath as a “man of the in-between” (a longstanding difficulty in translating 1 Samuel 17) appears to be a borrowing from Greek “man of the metaikhmion”, i.e., the space between two opposite army camps where champion combat would take place. Other scholars argue the description is a trustworthy reflection of the armaments that a Philistine warrior would have worn in the tenth century BCE.

    A story very similar to that of David and Goliath appears in the Iliad, written c.760–710 BCE, where the young Nestor fights and conquers the giant Ereuthalion. Each giant wields a distinctive weapon—an iron club in Ereuthalion’s case, a massive bronze spear in Goliath’s; each giant, clad in armor, comes out of the enemy’s massed array to challenge all the warriors in the opposing army; in each case the seasoned warriors are afraid, and the challenge is taken up by a stripling, the youngest in his family (Nestor is the twelfth son of Neleus, David the seventh or eighth son of Jesse). In each case an older and more experienced father figure (Nestor’s own father, David’s patron Saul) tells the boy that he is too young and inexperienced, but in each case the young hero receives divine aid and the giant is left sprawling on the ground. Nestor, fighting on foot, then takes the chariot of his enemy, while David, on foot, takes the sword of Goliath. The enemy army then flees, the victors pursue and slaughter them and return with their bodies, and the boy-hero is acclaimed by the people. However, some scholars question whether the biblical writers would have ever had access to the Iliad, and argue that the similarities between both tales are also present in other ancient Near Eastern accounts of duels.

    Goliath’s name

    Tell es-Safi, the biblical Gath and traditional home of Goliath, has been the subject of extensive excavations by Israel’s Bar-Ilan University. The archaeologists have established that this was one of the largest of the Philistine cities until it was destroyed in the ninth century BC, an event from which it never recovered. The Tell es-Safi inscription, a potsherd discovered at the site, and reliably dated to between the tenth to mid-ninth centuries BC, is inscribed with the two names ʾLWT and WLT. While the names are not directly connected with the biblical Goliath , they are etymologically related and demonstrate that the name fits with the context of the late tenth- to early ninth-century BC Philistine culture. The name “Goliath” itself is non-Semitic and has been linked with the Lydian king Alyattes, which also fits the Philistine context of the biblical Goliath story.A similar name, Uliat, is also attested in Carian inscriptions.[29] Aren Maeir, director of the excavation, comments: “Here we have very nice evidence the name Goliath appearing in the Bible in the context of the story of David and Goliath… is not some later literary creation.”

    Based on the southwest Anatolian onomastic considerations, Roger D. Woodard proposed *Walwatta as a reconstruction of the form ancestral to both Hebrew Goliath and Lydian Alyattes. In this case, the original meaning of Goliath’s name would be “Lion-man,” thus placing him within the realm of Indo-European warrior-beast mythology.

    The Babylonian Talmud explains the name “Goliath, son of Gath” through a reference to his mother’s promiscuity, based on the Aramaic גַּת (gat, winepress), as everyone threshed his mother as people do to grapes in a winepress (Sotah, 42b).

    The name sometimes appears in English as Goliah.

    Later traditions

    Judaism

    According to the Babylonian Talmud (Sotah 42b), Goliath was a son of Orpah, the sister-in-law of Ruth, David’s own great-grandmother (Ruth → Obed → Jesse → David). Ruth Rabbah, a haggadic and homiletic interpretation of the Book of Ruth, makes the blood relationship even closer, considering Orpah and Ruth to have been full sisters. Orpah was said to have made a pretense of accompanying Ruth but after forty paces left her. Thereafter she led a dissolute life. According to the Jerusalem Talmud, Goliath was born by polyspermy, and had about one hundred fathers.

    The Talmud stresses Goliath’s ungodliness: his taunts before the Israelites included the boast that it was he who had captured the Ark of the Covenant and brought it to the temple of Dagon, and his challenges to combat were made at morning and evening to disturb the Israelites in their prayers. His armor weighed 60 tons, according to rabbi Hanina; 120, according to rabbi Abba bar Kahana; and his sword, which became the sword of David, had marvelous powers. On his death it was found that his heart carried the image of Dagon, who thereby also came to a shameful downfall.

    In Pseudo-Philo, believed to have been composed between 135 BCE and 70 CE, David picks up seven stones and writes on them his father’s name, his own name, and the name of God, one name per stone; then, speaking to Goliath, he says:

    “Hear this word before you die: were not the two woman from whom you and I were born, sisters? And your mother was Orpah and my mother Ruth …”

    After David strikes Goliath with the stone he runs to Goliath before he dies, and Goliath says: “Hurry and kill me and rejoice.” David replies: “Before you die, open your eyes and see your slayer.” Goliath sees an angel and tells David that it is not he who has killed him but the angel. Pseudo-Philo then goes on to say that the angel of the Lord changes David’s appearance so that no one recognizes him, and thus Saul asks who he is.

    Islam

    Goliath appears in chapter 2 of the Quran (2: 247–252), in the narrative of David and Saul’s battle against the Philistines.Called Jalut in Arabic (جالوت), Goliath’s mention in the Quran is concise, although it remains a parallel to the account in the Hebrew Bible. Muslim scholars have tried to trace Goliath’s origins, most commonly with the Amalekites. Goliath, in early scholarly tradition, became a kind of byword or collective name for the oppressors of the Israelite nation before David. Muslim tradition sees the battle with Goliath as a prefiguration of Muhammad’s battle of Badr, and sees Goliath as parallel to the enemies that Muhammad faced.

    Modern usages of David and Goliath

    In modern usage, the phrase “David and Goliath” has taken on a secular meaning, denoting an underdog situation, a contest where a smaller, weaker opponent faces a much bigger, stronger adversary; if successful, the underdog may win in an unusual or surprising way.

    Theology professor Leonard Greenspoon, in his essay, “David vs. Goliath in the Sports Pages”, explains that “most writers use the story for its underdog overtones (the little guy wins) … Less likely to show up in newsprint is the contrast that was most important to the biblical authors: David’s victory shows the power of his God, while Goliath’s defeat reveals the weakness of the Philistine deities.”

    The phrase is widely used in news media to succinctly characterize underdog situations in many contexts without religious overtones. Contemporary headlines include: sports (“Haye relishes underdog role in ‘David and Goliath’ fight with Nikolai Valuev”—The Guardian); business (“On Internet, David-and-Goliath Battle Over Instant Messages”—The New York Times); science (“David and Goliath: How a tiny spider catches much larger prey”—ScienceDaily; politics (“Dissent in Cuba: David and Goliath”—The Economist); social justice (“David-and-Goliath Saga Brings Cable to Skid Row”—Los Angeles Times).

    Aside from the above allegorical use of “David and Goliath”, there is also the use of “Goliath” for a particularly tall person. For example, basketball player Wilt Chamberlain was nicknamed “Goliath”, which he disliked.

    American actor Ted Cassidy portrayed Goliath in the TV series Greatest Heroes of the Bible (1978). Italian actor Luigi Montefiori portrayed this 9 ft 0 in (2.74 m)-tall giant in Paramount’s 1985 live-action film King David as part of a flashback. This film includes the King of the Philistines saying: “Goliath has challenged the Israelites six times and no one has responded.” It is then on the seventh time that David meets his challenge.

    Toho and Tsuburaya Productions collaborated on a film called Daigoro vs. Goliath (1972), which follows the story relatively closely but recasts the main characters as kaiju.

    In 2005, Lightstone Studios released a direct-to-DVD movie musical titled “One Smooth Stone”, which was later changed to “David and Goliath”. It is part of the Liken the scriptures (now just Liken) series of movie musicals on DVD based on scripture stories. Thurl Bailey, a former NBA basketball player, was cast to play the part of Goliath in this film.

    In 2009, NBC aired Kings, which has a narrative loosely based on the biblical story of King David, but set in a kingdom that culturally and technologically resembles the present-day United States. The part of Goliath is portrayed by a tank, which David destroys with a shoulder-fired rocket launcher.

    In 1975, Kaveret recorded and released a humorous interpretation of the Goliath story, with several changes made such as Goliath being the “Demon from Ashkelon”, and David randomly meeting Goliath rather than dueling each other on a battlefield.

    Italian Goliath film series (1960–1964)

    The Italians used Goliath as an action superhero in a series of biblical adventure films (peplums) in the early 1960s. He possessed amazing strength, and the films were similar in theme to their Hercules and Maciste movies. After the classic Hercules (1958) became a blockbuster sensation in the film industry, the 1959 Steve Reeves film Terrore dei Barbari (Terror of the Barbarians) was retitled Goliath and the Barbarians in the United States, (after Joseph E. Levine claimed the sole right to the name of Hercules); the film was so successful at the box office, it inspired Italian filmmakers to do a series of four more films featuring a beefcake hero named Goliath, although the films were not really related to each other. Note that the Italian film David and Goliath (1960), starring Orson Welles, was not one of these, since that film was a straightforward adaptation of the biblical story.

    The four titles in the Italian Goliath series were as follows:

    Goliath contro i giganti/Goliath Against the Giants (1960) starring Brad Harris
    Goliath e la schiava ribelle/Goliath and the Rebel Slave (a.k.a. The Tyrant of Lydia vs. The Son of Hercules) (1963) starring Gordon Scott
    Golia e il cavaliere mascherato/Goliath and the Masked Rider (a.k.a. Hercules and the Masked Rider) (1964) starring Alan Steel
    Golia alla conquista di Bagdad/Goliath at the Conquest of Baghdad (a.k.a. Goliath at the Conquest of Damascus, 1964) starring Peter Lupus


    The name Goliath was later inserted into the film titles of three other Italian muscle man movies that were retitled for distribution in the United States in an attempt to cash in on the Goliath craze, but these films were not originally made as Goliath films in Italy.

    Both Goliath and the Vampires (1961) and Goliath and the Sins of Babylon (1963) actually featured the famed superhero Maciste in the original Italian versions, but American distributors did not feel the name Maciste had any meaning to American audiences. Goliath and the Dragon (1960) was originally an Italian Hercules film called The Revenge of Hercules.

  • Proposition

    A proposition is a central concept in the philosophy of language, semantics, logic, and related fields, often characterized as the primary bearer of truth or falsity. Propositions are also often characterized as the type of object that declarative sentences denote. For instance, the sentence “The sky is blue” denotes the proposition that the sky is blue. However, crucially, propositions are not themselves linguistic expressions. For instance, the English sentence “Snow is white” denotes the same proposition as the German sentence “Schnee ist weiß” even though the two sentences are not the same. Similarly, propositions can also be characterized as the objects of belief and other propositional attitudes. For instance if someone believes that the sky is blue, the object of their belief is the proposition that the sky is blue.

    Formally, propositions are often modeled as functions which map a possible world to a truth value. For instance, the proposition that the sky is blue can be modeled as a function which would return the truth value T if given the actual world as input, but would return F if given some alternate world where the sky is green. However, a number of alternative formalizations have been proposed, notably the structured propositions view.

    Propositions have played a large role throughout the history of logic, linguistics, philosophy of language, and related disciplines. Some researchers have doubted whether a consistent definition of propositionhood is possible, David Lewis even remarking that “the conception we associate with the word ‘proposition’ may be something of a jumble of conflicting desiderata”. The term is often used broadly and has been used to refer to various related concepts.

    Relation to the mind

    In relation to the mind, propositions are discussed primarily as they fit into propositional attitudes. Propositional attitudes are simply attitudes characteristic of folk psychology (belief, desire, etc.) that one can take toward a proposition (e.g. ‘it is raining,’ ‘snow is white,’ etc.). In English, propositions usually follow folk psychological attitudes by a “that clause” (e.g. “Jane believes that it is raining”). In philosophy of mind and psychology, mental states are often taken to primarily consist in propositional attitudes. The propositions are usually said to be the “mental content” of the attitude. For example, if Jane has a mental state of believing that it is raining, her mental content is the proposition ‘it is raining.’ Furthermore, since such mental states are about something (namely, propositions), they are said to be intentional mental states.

    Explaining the relation of propositions to the mind is especially difficult for non-mentalist views of propositions, such as those of the logical positivists and Russell described above, and Gottlob Frege’s view that propositions are Platonist entities, that is, existing in an abstract, non-physical realm. So some recent views of propositions have taken them to be mental. Although propositions cannot be particular thoughts since those are not shareable, they could be types of cognitive events or properties of thoughts (which could be the same across different thinkers).

    Philosophical debates surrounding propositions as they relate to propositional attitudes have also recently centered on whether they are internal or external to the agent, or whether they are mind-dependent or mind-independent entities. For more, see the entry on internalism and externalism in philosophy of mind.

    Objections to propositions

    Attempts to provide a workable definition of proposition include the following:

    Two meaningful declarative sentences express the same proposition, if and only if they mean the same thing.

    which defines proposition in terms of synonymity. For example, “Snow is white” (in English) and “Schnee ist weiß” (in German) are different sentences, but they say the same thing, so they express the same proposition. Another definition of proposition is:

    Two meaningful declarative sentence-tokens express the same proposition, if and only if they mean the same thing.

    The above definitions can result in two identical sentences/sentence-tokens appearing to have the same meaning, and thus expressing the same proposition and yet having different truth-values, as in “I am Spartacus” said by Spartacus and said by John Smith, and “It is Wednesday” said on a Wednesday and on a Thursday. These examples reflect the problem of ambiguity in common language, resulting in a mistaken equivalence of the statements. “I am Spartacus” spoken by Spartacus is the declaration that the individual speaking is called Spartacus and it is true. When spoken by John Smith, it is a declaration about a different speaker and it is false. The term “I” means different things, so “I am Spartacus” means different things.

    A related problem is when identical sentences have the same truth-value, yet express different propositions. The sentence “I am a philosopher” could have been spoken by both Socrates and Plato. In both instances, the statement is true, but means something different.

    These problems are addressed in predicate logic by using a variable for the problematic term, so that “X is a philosopher” can have Socrates or Plato substituted for X, illustrating that “Socrates is a philosopher” and “Plato is a philosopher” are different propositions. Similarly, “I am Spartacus” becomes “X is Spartacus”, where X is replaced with terms representing the individuals Spartacus and John Smith.

    In other words, the example problems can be averted if sentences are formulated with precision such that their terms have unambiguous meanings.

    A number of philosophers and linguists claim that all definitions of a proposition are too vague to be useful. For them, it is just a misleading concept that should be removed from philosophy and semantics. W. V. Quine, who granted the existence of sets in mathematics, maintained that the indeterminacy of translation prevented any meaningful discussion of propositions, and that they should be discarded in favor of sentences. P. F. Strawson, on the other hand, advocated for the use of the term “statement”.

    Historical usage

    By Aristotle

    In Aristotelian logic a proposition was defined as a particular kind of sentence (a declarative sentence) that affirms or denies a predicate of a subject, optionally with the help of a copula. Aristotelian propositions take forms like “All men are mortal” and “Socrates is a man.”

    Aristotelian logic identifies a categorical proposition as a sentence which affirms or denies a predicate of a subject, optionally with the help of a copula. An Aristotelian proposition may take the form of “All men are mortal” or “Socrates is a man.” In the first example, the subject is “men”, predicate is “mortal” and copula is “are”, while in the second example, the subject is “Socrates”, the predicate is “a man” and copula is “is”.

    By the logical positivists

    Often, propositions are related to closed formulae (or logical sentence) to distinguish them from what is expressed by an open formula. In this sense, propositions are “statements” that are truth-bearers. This conception of a proposition was supported by the philosophical school of logical positivism.

    Some philosophers argue that some (or all) kinds of speech or actions besides the declarative ones also have propositional content. For example, yes–no questions present propositions, being inquiries into the truth value of them. On the other hand, some signs can be declarative assertions of propositions, without forming a sentence nor even being linguistic (e.g. traffic signs convey definite meaning which is either true or false).

    Propositions are also spoken of as the content of beliefs and similar intentional attitudes, such as desires, preferences, and hopes. For example, “I desire that I have a new car”, or “I wonder whether it will snow” (or, whether it is the case that “it will snow”). Desire, belief, doubt, and so on, are thus called propositional attitudes when they take this sort of content.

    By Russell

    Bertrand Russell held that propositions were structured entities with objects and properties as constituents. One important difference between Ludwig Wittgenstein’s view (according to which a proposition is the set of possible worlds/states of affairs in which it is true) is that on the Russellian account, two propositions that are true in all the same states of affairs can still be differentiated. For instance, the proposition “two plus two equals four” is distinct on a Russellian account from the proposition “three plus three equals six”. If propositions are sets of possible worlds, however, then all mathematical truths (and all other necessary truths) are the same set (the set of all possible worlds).

  • Evidence

    Evidence for a proposition is what supports the proposition. It is usually understood as an indication that the proposition is true. The exact definition and role of evidence vary across different fields. In epistemology, evidence is what justifies beliefs or what makes it rational to hold a certain doxastic attitude. For example, a perceptual experience of a tree may serve as evidence to justify the belief that there is a tree. In this role, evidence is usually understood as a private mental state. In phenomenology, evidence is limited to intuitive knowledge, often associated with the controversial assumption that it provides indubitable access to truth.

    In the science, scientific evidence is information gained through the scientific method that confirms or disconfirms scientific hypotheses, acting as a neutral arbiter between competing theories. Measurements of Mercury’s “anomalous” orbit, for example, are seen as evidence that confirms Einstein’s theory of general relativity. The problems of underdetermination and theory-ladenness are two obstacles that threaten to undermine the role of scientific evidence. Philosophers of science tend to understand evidence not as mental states but as verifiable information, observable physical objects or events, secured by following the scientific method.

    In law, evidence is information to establish or refute claims relevant to a case, such as testimony, documentary evidence, and physical evidence.

    The relation between evidence and a supported statement can vary in strength, ranging from weak correlation to indisputable proof. Theories of the evidential relation examine the nature of this connection. Probabilistic approaches hold that something counts as evidence if it increases the probability of the supported statement. According to hypothetico-deductivism, evidence consists in observational consequences of a hypothesis. The positive-instance approach states that an observation sentence is evidence for a universal statement if the sentence describes a positive instance of this statement.

    Nature of evidence

    Notion

    Understood in its broadest sense, evidence for a proposition is what supports this proposition. Traditionally, the term is sometimes understood in a narrower sense: as the intuitive knowledge of facts that are considered indubitable. In this sense, only the singular form is used. This meaning is found especially in phenomenology, in which evidence is elevated to one of the basic principles of philosophy, giving philosophy the ultimate justifications that are supposed to turn it into a rigorous science. In a more modern usage, the plural form is also used. In academic discourse, evidence plays a central role in epistemology and in the philosophy of science. Reference to evidence is made in many different fields, like in science, in the legal system, in history, in journalism and in everyday discourse. A variety of different attempts have been made to conceptualize the nature of evidence. These attempts often proceed by starting with intuitions from one field or in relation to one theoretical role played by evidence and go on to generalize these intuitions, leading to a universal definition of evidence.

    One important intuition is that evidence is what justifies beliefs. This line of thought is usually followed in epistemology and tends to explain evidence in terms of private mental states, for example, as experiences, other beliefs or knowledge. This is closely related to the idea that how rational someone is, is determined by how they respond to evidence.Another intuition, which is more dominant in the philosophy of science, focuses on evidence as that which confirms scientific hypotheses and arbitrates between competing theories. On this view, it is essential that evidence is public so that different scientists can share the same evidence. This leaves publicly observable phenomena like physical objects and events as the best candidates for evidence, unlike private mental states. One problem with these approaches is that the resulting definitions of evidence, both within a field and between fields, vary a lot and are incompatible with each other. For example, it is not clear what a bloody knife and a perceptual experience have in common when both are treated as evidence in different disciplines. This suggests that there is no unitary concept corresponding to the different theoretical roles ascribed to evidence, i.e. that we do not always mean the same thing when we talk of evidence.

    Characteristics

    On the other hand, Aristotle, phenomenologists, and numerous scholars accept that there could be several degrees of evidence. For instance, while the outcome of a complex equation may become more or less evident to a mathematician after hours of deduction, yet with little doubts about it, a simpler formula would appear more evident to them.

    Riofrio has detected some characteristics that are present in evident arguments and proofs. The more they are evident, the more these characteristics will be present. There are six intrinsic characteristics of evidence:

    The truth lies in what is evident, while falsehood or irrationality, although it may appear evident at times, lacks true evidence.
    What is evident aligns coherently with other truths acquired through knowledge. Any insurmountable incoherence would indicate the presence of error or falsehood.
    Evident truths are based on necessary reasoning.
    The simplest truths are the most evident. They are self-explanatory and do not require argumentation to be understood by the intellect. However, for those lacking education, certain complex truths require rational discourse to become evident.
    Evident truths do not need justification; they are indubitable. They are intuitively grasped by the intellect, without the need for further discourse, arguments, or proof.
    Evident truths are clear, translucent, and filled with light.
    In addition, four subjective or external characteristics can be detected over those things that are more or less evident:

    The evident instills certainty and grants the knower a subjective sense of security, as they believe to have aligned with the truth
    Initially, evident truths are perceived as natural and effortless, as Aristotle highlighted. They are innately present within the intellect, fostering a peaceful and harmonious understanding.
    Consequently, evident truths appear to be widely shared, strongly connected to common sense, which comprises generally accepted beliefs.
    Evident truths are fertile ground: they provide a solid foundation for other branches of scientific knowledge to flourish.
    These ten characteristics of what is evident allowed Riofrio to formulate a test of evidence to detect the level of certainty or evidence that one argument or proof could have.

    Different approaches to evidence

    Important theorists of evidence include Bertrand Russell, Willard Van Orman Quine, the logical positivists, Timothy Williamson, Earl Conee and Richard Feldman. Russell, Quine and the logical positivists belong to the empiricist tradition and hold that evidence consists in sense data, stimulation of one’s sensory receptors and observation statements, respectively. According to Williamson, all and only knowledge constitute evidence.Conee and Feldman hold that only one’s current mental states should be considered evidence.

    In epistemology

    The guiding intuition within epistemology concerning the role of evidence is that it is what justifies beliefs. For example, Phoebe’s auditory experience of the music justifies her belief that the speakers are on. Evidence has to be possessed by the believer in order to play this role. So Phoebe’s own experiences can justify her own beliefs but not someone else’s beliefs. Some philosophers hold that evidence possession is restricted to conscious mental states, for example, to sense data. This view has the implausible consequence that many of simple everyday-beliefs would be unjustified. The more common view is that all kinds of mental states, including stored beliefs that are currently unconscious, can act as evidence. It is sometimes argued that the possession of a mental state capable of justifying another is not sufficient for the justification to happen. The idea behind this line of thought is that justified belief has to be connected to or grounded in the mental state acting as its evidence. So Phoebe’s belief that the speakers are on is not justified by her auditory experience if the belief is not based in this experience. This would be the case, for example, if Phoebe has both the experience and the belief but is unaware of the fact that the music is produced by the speakers.

    It is sometimes held that only propositional mental states can play this role, a position known as “propositionalism”. A mental state is propositional if it is an attitude directed at a propositional content. Such attitudes are usually expressed by verbs like “believe” together with a that-clause, as in “Robert believes that the corner shop sells milk”.Such a view denies that sensory impressions can act as evidence. This is often held as an argument against this view since sensory impressions are commonly treated as evidence. Propositionalism is sometimes combined with the view that only attitudes to true propositions can count as evidence. On this view, the belief that the corner shop sells milk only constitutes evidence for the belief that the corner shop sells dairy products if the corner shop actually sells milk. Against this position, it has been argued that evidence can be misleading but still count as evidence.

    This line of thought is often combined with the idea that evidence, propositional or otherwise, determines what it is rational for us to believe. But it can be rational to have a false belief. This is the case when we possess misleading evidence. For example, it was rational for Neo in the Matrix movie to believe that he was living in the 20th century because of all the evidence supporting his belief despite the fact that this evidence was misleading since it was part of a simulated reality. This account of evidence and rationality can also be extended to other doxastic attitudes, like disbelief and suspension of belief. So rationality does not just demand that we believe something if we have decisive evidence for it, it also demands that we disbelieve something if we have decisive evidence against it and that we suspend belief if we lack decisive evidence either way.

    In phenomenology

    The meaning of the term “evidence” in phenomenology shows many parallels to its epistemological usage, but it is understood in a narrower sense. Thus, evidence here specifically refers to intuitive knowledge, which is described as “self-given” (selbst-gegeben). This contrasts with empty intentions, in which one refers to states of affairs through a certain opinion, but without an intuitive presentation. This is why evidence is often associated with the controversial thesis that it constitutes an immediate access to truth. In this sense, the evidently given phenomenon guarantees its own truth and is therefore considered indubitable. Due to this special epistemological status of evidence, it is regarded in phenomenology as the basic principle of all philosophy.In this form, it represents the lowest foundation of knowledge, which consists of indubitable insights upon which all subsequent knowledge is built. This evidence-based method is meant to make it possible for philosophy to overcome many of the traditionally unresolved disagreements and thus become a rigorous science.This far-reaching claim of phenomenology, based on absolute certainty, is one of the focal points of criticism by its opponents. Thus, it has been argued that even knowledge based on self-evident intuition is fallible. This can be seen, for example, in the fact that even among phenomenologists, there is much disagreement about the basic structures of experience.

    In science

    In the sciences, evidence is understood as what confirms or disconfirms scientific hypotheses. The term “confirmation” is sometimes used synonymously with that of “evidential support”. Measurements of Mercury’s “anomalous” orbit, for example, are seen as evidence that confirms Einstein’s theory of general relativity. This is especially relevant for choosing between competing theories. So in the case above, evidence plays the role of neutral arbiter between Newton’s and Einstein’s theory of gravitation. This is only possible if scientific evidence is public and uncontroversial so that proponents of competing scientific theories agree on what evidence is available. These requirements suggest scientific evidence consists not of private mental states but of public physical objects or events.

    It is often held that evidence is in some sense prior to the hypotheses it confirms. This was sometimes understood as temporal priority, i.e. that we come first to possess the evidence and later form the hypothesis through induction. But this temporal order is not always reflected in scientific practice, where experimental researchers may look for a specific piece of evidence in order to confirm or disconfirm a pre-existing hypothesis. Logical positivists, on the other hand, held that this priority is semantic in nature, i.e. that the meanings of the theoretical terms used in the hypothesis are determined by what would count as evidence for them. Counterexamples for this view come from the fact that our idea of what counts as evidence may change while the meanings of the corresponding theoretical terms remain constant. The most plausible view is that this priority is epistemic in nature, i.e. that our belief in a hypothesis is justified based on the evidence while the justification for the belief in the evidence does not depend on the hypothesis.

    A central issue for the scientific conception of evidence is the problem of underdetermination, i.e. that the evidence available supports competing theories equally well. So, for example, evidence from our everyday life about how gravity works confirms Newton’s and Einstein’s theory of gravitation equally well and is therefore unable to establish consensus among scientists. But in such cases, it is often the gradual accumulation of evidence that eventually leads to an emerging consensus. This evidence-driven process towards consensus seems to be one hallmark of the sciences not shared by other fields.

    Another problem for the conception of evidence in terms of confirmation of hypotheses is that what some scientists consider the evidence to be may already involve various theoretical assumptions not shared by other scientists. This phenomenon is known as theory-ladenness. Some cases of theory-ladenness are relatively uncontroversial, for example, that the numbers output by a measurement device need additional assumptions about how this device works and what was measured in order to count as meaningful evidence.Other putative cases are more controversial, for example, the idea that different people or cultures perceive the world through different, incommensurable conceptual schemes, leading them to very different impressions about what is the case and what evidence is available. Theory-ladenness threatens to impede the role of evidence as neutral arbiter since these additional assumptions may favor some theories over others. It could thereby also undermine a consensus to emerge since the different parties may be unable to agree even on what the evidence is. When understood in the widest sense, it is not controversial that some form of theory-ladenness exists. But it is questionable whether it constitutes a serious threat to scientific evidence when understood in this sense.

    Nature of the evidential relation

    Philosophers in the 20th century started to investigate the “evidential relation”, the relation between evidence and the proposition supported by it. The issue of the nature of the evidential relation concerns the question of what this relation has to be like in order for one thing to justify a belief or to confirm a hypothesis. Important theories in this field include the probabilistic approach, hypothetico-deductivism and the positive-instance approach.

    Probabilistic approaches, also referred to as Bayesian confirmation theory, explain the evidential relation in terms of probabilities. They hold that all that is necessary is that the existence of the evidence increases the likelihood that the hypothesis is true. In words: a piece of evidence (E) confirms a hypothesis (H) if the conditional probability of this hypothesis relative to the evidence is higher than the unconditional probability of the hypothesis by itself. Smoke (E), for example, is evidence that there is a fire (H), because the two usually occur together, which is why the likelihood of fire given that there is smoke is higher than the likelihood of fire by itself. On this view, evidence is akin to an indicator or a symptom of the truth of the hypothesis. Against this approach, it has been argued that it is too liberal because it allows accidental generalizations as evidence. Finding a nickel in one’s pocket, for example, raises the probability of the hypothesis that “All the coins in my pockets are nickels”. But, according to Alvin Goldman, it should not be considered evidence for this hypothesis since there is no lawful connection between this one nickel and the other coins in the pocket.

    Hypothetico-deductivism is a non-probabilistic approach that characterizes the evidential relations in terms of deductive consequences of the hypothesis. According to this view, ”evidence for a hypothesis is a true observational consequence of that hypothesis”. One problem with the characterization so far is that hypotheses usually contain relatively little information and therefore have few if any deductive observational consequences. So the hypothesis by itself that there is a fire does not entail that smoke is observed. Instead, various auxiliary assumptions have to be included about the location of the smoke, the fire, the observer, the lighting conditions, the laws of chemistry, etc. In this way, the evidential relation becomes a three-place relation between evidence, hypothesis and auxiliary assumptions. This means that whether a thing is evidence for a hypothesis depends on the auxiliary assumptions one holds. This approach fits well with various scientific practices. For example, it is often the case that experimental scientists try to find evidence that would confirm or disconfirm a proposed theory. The hypothetico-deductive approach can be used to predict what should be observed in an experiment if the theory was true.It thereby explains the evidential relation between the experiment and the theory.One problem with this approach is that it cannot distinguish between relevant and certain irrelevant cases. So if smoke is evidence for the hypothesis “there is fire”, then it is also evidence for conjunctions including this hypothesis, for example, “there is fire and Socrates was wise”, despite the fact that Socrates’s wisdom is irrelevant here.

    According to the positive-instance approach, an observation sentence is evidence for a universal hypothesis if the sentence describes a positive instance of this hypothesis. For example, the observation that “this swan is white” is an instance of the universal hypothesis that “all swans are white”. This approach can be given a precise formulation in first-order logic: a proposition is evidence for a hypothesis if it entails the “development of the hypothesis”. Intuitively, the development of the hypothesis is what the hypothesis states if it was restricted to only the individuals mentioned in the evidence. In the case above, we have the hypothesis which, when restricted to the domain “{a}”, containing only the one individual mentioned in the evidence, entails the evidence, i.e.One important shortcoming of this approach is that it requires that the hypothesis and the evidence are formulated in the same vocabulary, i.e. use the same predicates, above. But many scientific theories posit theoretical objects, like electrons or strings in physics, that are not directly observable and therefore cannot show up in the evidence as conceived here.

    Empirical evidence (in science)

    In scientific research evidence is accumulated through observations of phenomena that occur in the natural world, or which are created as experiments in a laboratory or other controlled conditions. Scientists tend to focus on how the data used during statistical inference are generated. Scientific evidence usually goes towards supporting or rejecting a hypothesis.

    The burden of proof is on the person making a contentious claim. Within science, this translates to the burden resting on presenters of a paper, in which the presenters argue for their specific findings. This paper is placed before a panel of judges where the presenter must defend the thesis against all challenges.

    When evidence is contradictory to predicted expectations, the evidence and the ways of making it are often closely scrutinized (see experimenter’s regress) and only at the end of this process is the hypothesis rejected: this can be referred to as ‘refutation of the hypothesis’. The rules for evidence used by science are collected systematically in an attempt to avoid the bias inherent to anecdotal evidence.

    Law

    In law, the production and presentation of evidence depend first on establishing on whom the burden of proof lies. Admissible evidence is that which a court receives and considers for the purposes of deciding a particular case. Two primary burden-of-proof considerations exist in law. The first is on whom the burden rests. In many, especially Western, courts, the burden of proof is placed on the prosecution in criminal cases and the plaintiff in civil cases. The second consideration is the degree of certitude proof must reach, depending on both the quantity and quality of evidence. These degrees are different for criminal and civil cases, the former requiring evidence beyond a reasonable doubt, the latter considering only which side has the preponderance of evidence, or whether the proposition is more likely true or false.

    The parts of a legal case that are not in controversy are known, in general, as the “facts of the case.” Beyond any facts that are undisputed, a judge or jury is usually tasked with being a trier of fact for the other issues of a case. Evidence and rules are used to decide questions of fact that are disputed, some of which may be determined by the legal burden of proof relevant to the case. Evidence in certain cases (e.g. capital crimes) must be more compelling than in other situations (e.g. minor civil disputes), which drastically affects the quality and quantity of evidence necessary to decide a case. The decision-maker, often a jury, but sometimes a judge decides whether the burden of proof has been fulfilled. After deciding who will carry the burden of proof, the evidence is first gathered and then presented before the court:

    Collection

    In a criminal investigation, rather than attempting to prove an abstract or hypothetical point, the evidence gatherers attempt to determine who is responsible for a criminal act. The focus of criminal evidence is to connect physical evidence and reports of witnesses to a specific person

    Presentation

    The path that physical evidence takes from the scene of a crime or the arrest of a suspect to the courtroom is called the chain of custody. In a criminal case, this path must be clearly documented or attested to by those who handled the evidence. If the chain of evidence is broken, a defendant may be able to persuade the judge to declare the evidence inadmissible.

    Presenting evidence before the court differs from the gathering of evidence in important ways. Gathering evidence may take many forms; presenting evidence that tends to prove or disprove the point at issue is strictly governed by rules. Failure to follow these rules leads to any number of consequences. In law, certain policies allow (or require) evidence to be excluded from consideration based either on indicia relating to reliability, or broader social concerns. Testimony (which tells) and exhibits (which show) are the two main categories of evidence presented at a trial or hearing. In the United States, evidence in federal court is admitted or excluded under the Federal Rules of Evidence.

    Burden of proof

    The burden of proof is the obligation of a party in an argument or dispute to provide sufficient evidence to shift the other party’s or a third party’s belief from their initial position. The burden of proof must be fulfilled by both establishing confirming evidence and negating oppositional evidence. Conclusions drawn from evidence may be subject to criticism based on a perceived failure to fulfill the burden of proof.

    Two principal considerations are:

    On whom does the burden of proof rest?
    To what degree of certitude must the assertion be supported?
    The latter question depends on the nature of the point under contention and determines the quantity and quality of evidence required to meet the burden of proof.

    In a criminal trial in the United States, for example, the prosecution carries the burden of proof since the defendant is presumed innocent until proven guilty beyond a reasonable doubt. Similarly, in most civil procedures, the plaintiff carries the burden of proof and must convince a judge or jury that the preponderance of the evidence is on their side. Other legal standards of proof include “reasonable suspicion”, “probable cause” (as for arrest), “prima facie evidence”, “credible evidence”, “substantial evidence”, and “clear and convincing evidence”.

    In a philosophical debate, there is an implicit burden of proof on the party asserting a claim, since the default position is generally one of neutrality or unbelief. Each party in a debate will therefore carry the burden of proof for any assertion they make in the argument, although some assertions may be granted by the other party without further evidence. If the debate is set up as a resolution to be supported by one side and refuted by another, the overall burden of proof is on the side supporting the resolution.

  • Empirical evidence

    Empirical evidence is evidence obtained through sense experience or experimental procedure. It is of central importance to the sciences and plays a role in various other fields, like epistemology and law.

    There is no general agreement on how the terms evidence and empirical are to be defined. Often different fields work with quite different conceptions. In epistemology, evidence is what justifies beliefs or what determines whether holding a certain belief is rational. This is only possible if the evidence is possessed by the person, which has prompted various epistemologists to conceive evidence as private mental states like experiences or other beliefs. In philosophy of science, on the other hand, evidence is understood as that which confirms or disconfirms scientific hypotheses and arbitrates between competing theories. For this role, evidence must be public and uncontroversial, like observable physical objects or events and unlike private mental states, so that evidence may foster scientific consensus. The term empirical comes from Greek ἐμπειρία empeiría, i.e. ‘experience’. In this context, it is usually understood as what is observable, in contrast to unobservable or theoretical objects. It is generally accepted that unaided perception constitutes observation, but it is disputed to what extent objects accessible only to aided perception, like bacteria seen through a microscope or positrons detected in a cloud chamber, should be regarded as observable.

    Empirical evidence is essential to a posteriori knowledge or empirical knowledge, knowledge whose justification or falsification depends on experience or experiment. A priori knowledge, on the other hand, is seen either as innate or as justified by rational intuition and therefore as not dependent on empirical evidence. Rationalism fully accepts that there is knowledge a priori, which is either outright rejected by empiricism or accepted only in a restricted way as knowledge of relations between our concepts but not as pertaining to the external world.

    Scientific evidence is closely related to empirical evidence but not all forms of empirical evidence meet the standards dictated by scientific methods. Sources of empirical evidence are sometimes divided into observation and experimentation, the difference being that only experimentation involves manipulation or intervention: phenomena are actively created instead of being passively observed.

    Background

    The concept of evidence is of central importance in epistemology and in philosophy of science but plays different roles in these two fields. In epistemology, evidence is what justifies beliefs or what determines whether holding a certain doxastic attitude is rational. For example, the olfactory experience of smelling smoke justifies or makes it rational to hold the belief that something is burning. It is usually held that for justification to work, the evidence has to be possessed by the believer. The most straightforward way to account for this type of evidence possession is to hold that evidence consists of the private mental states possessed by the believer.

    Some philosophers restrict evidence even further, for example, to only conscious, propositional or factive mental states. Restricting evidence to conscious mental states has the implausible consequence that many simple everyday beliefs would be unjustified. This is why it is more common to hold that all kinds of mental states, including stored but currently unconscious beliefs, can act as evidence.Various of the roles played by evidence in reasoning, for example, in explanatory, probabilistic and deductive reasoning, suggest that evidence has to be propositional in nature, i.e. that it is correctly expressed by propositional attitude verbs like “believe” together with a that-clause, like “that something is burning”. But it runs counter to the common practice of treating non-propositional sense-experiences, like bodily pains, as evidence. Its defenders sometimes combine it with the view that evidence has to be factive, i.e. that only attitudes towards true propositions constitute evidence. In this view, there is no misleading evidence. The olfactory experience of smoke would count as evidence if it was produced by a fire but not if it was produced by a smoke generator. This position has problems in explaining why it is still rational for the subject to believe that there is a fire even though the olfactory experience cannot be considered evidence.

    In philosophy of science, evidence is understood as that which confirms or disconfirms scientific hypotheses and arbitrates between competing theories. Measurements of Mercury’s “anomalous” orbit, for example, constitute evidence that plays the role of neutral arbiter between Newton’s and Einstein’s theory of gravitation by confirming Einstein’s theory. For scientific consensus, it is central that evidence is public and uncontroversial, like observable physical objects or events and unlike private mental states.This way it can act as a shared ground for proponents of competing theories. Two issues threatening this role are the problem of underdetermination and theory-ladenness. The problem of underdetermination concerns the fact that the available evidence often provides equal support to either theory and therefore cannot arbitrate between them.Theory-ladenness refers to the idea that evidence already includes theoretical assumptions. These assumptions can hinder it from acting as neutral arbiter. It can also lead to a lack of shared evidence if different scientists do not share these assumptions. Thomas Kuhn is an important advocate of the position that theory-ladenness concerning scientific paradigms plays a central role in science.

    Definition

    A thing is evidence for a proposition if it epistemically supports this proposition or indicates that the supported proposition is true. Evidence is empirical if it is constituted by or accessible to sensory experience. There are various competing theories about the exact definition of the terms evidence and empirical. Different fields, like epistemology, the sciences or legal systems, often associate different concepts with these terms. An important distinction among theories of evidence is whether they identify evidence with private mental states or with public physical objects. Concerning the term empirical, there is a dispute about where to draw the line between observable or empirical objects in contrast to unobservable or merely theoretical objects.

    The traditional view proposes that evidence is empirical if it is constituted by or accessible to sensory experience. This involves experiences arising from the stimulation of the sense organs, like visual or auditory experiences, but the term is often used in a wider sense including memories and introspection. It is usually seen as excluding purely intellectual experiences, like rational insights or intuitions used to justify basic logical or mathematical principles.The terms empirical and observable are closely related and sometimes used as synonyms.

    There is an active debate in contemporary philosophy of science as to what should be regarded as observable or empirical in contrast to unobservable or merely theoretical objects. There is general consensus that everyday objects like books or houses are observable since they are accessible via unaided perception, but disagreement starts for objects that are only accessible through aided perception. This includes using telescopes to study distant galaxies, microscopes to study bacteria or using cloud chambers to study positrons.So the question is whether distant galaxies, bacteria or positrons should be regarded as observable or merely theoretical objects. Some even hold that any measurement process of an entity should be considered an observation of this entity. In this sense, the interior of the Sun is observable since neutrinos originating there can be detected. The difficulty with this debate is that there is a continuity of cases going from looking at something with the naked eye, through a window, through a pair of glasses, through a microscope, etc. Because of this continuity, drawing the line between any two adjacent cases seems to be arbitrary. One way to avoid these difficulties is to hold that it is a mistake to identify the empirical with what is observable or sensible. Instead, it has been suggested that empirical evidence can include unobservable entities as long as they are detectable through suitable measurements. A problem with this approach is that it is rather far from the original meaning of “empirical”, which contains the reference to experience.

    Related concepts

    Knowledge a posteriori and a priori

    Knowledge or the justification of a belief is said to be a posteriori if it is based on empirical evidence. A posteriori refers to what depends on experience (what comes after experience), in contrast to a priori, which stands for what is independent of experience (what comes before experience). For example, the proposition that “all bachelors are unmarried” is knowable a priori since its truth only depends on the meanings of the words used in the expression. The proposition “some bachelors are happy”, on the other hand, is only knowable a posteriori since it depends on experience of the world as its justifier. Immanuel Kant held that the difference between a posteriori and a priori is tantamount to the distinction between empirical and non-empirical knowledge.

    Two central questions for this distinction concern the relevant sense of “experience” and of “dependence”. The paradigmatic justification of knowledge a posteriori consists in sensory experience, but other mental phenomena, like memory or introspection, are also usually included in it. But purely intellectual experiences, like rational insights or intuitions used to justify basic logical or mathematical principles, are normally excluded from it. There are different senses in which knowledge may be said to depend on experience. In order to know a proposition, the subject has to be able to entertain this proposition, i.e. possess the relevant concepts. For example, experience is necessary to entertain the proposition “if something is red all over then it is not green all over” because the terms “red” and “green” have to be acquired this way. But the sense of dependence most relevant to empirical evidence concerns the status of justification of a belief. So experience may be needed to acquire the relevant concepts in the example above, but once these concepts are possessed, no further experience providing empirical evidence is needed to know that the proposition is true, which is why it is considered to be justified a priori.

    Empiricism and rationalism

    In its strictest sense, empiricism is the view that all knowledge is based on experience or that all epistemic justification arises from empirical evidence. This stands in contrast to the rationalist view, which holds that some knowledge is independent of experience, either because it is innate or because it is justified by reason or rational reflection alone. Expressed through the distinction between knowledge a priori and a posteriori from the previous section, rationalism affirms that there is knowledge a priori, which is denied by empiricism in this strict form. One difficulty for empiricists is to account for the justification of knowledge pertaining to fields like mathematics and logic, for example, that 3 is a prime number or that modus ponens is a valid form of deduction. The difficulty is due to the fact that there seems to be no good candidate of empirical evidence that could justify these beliefs. Such cases have prompted empiricists to allow for certain forms of knowledge a priori, for example, concerning tautologies or relations between our concepts. These concessions preserve the spirit of empiricism insofar as the restriction to experience still applies to knowledge about the external world. In some fields, like metaphysics or ethics, the choice between empiricism and rationalism makes a difference not just for how a given claim is justified but for whether it is justified at all. This is best exemplified in metaphysics, where empiricists tend to take a skeptical position, thereby denying the existence of metaphysical knowledge, while rationalists seek justification for metaphysical claims in metaphysical intuitions.

    Scientific evidence

    Scientific evidence is closely related to empirical evidence. Some theorists, like Carlos Santana, have argued that there is a sense in which not all empirical evidence constitutes scientific evidence. One reason for this is that the standards or criteria that scientists apply to evidence exclude certain evidence that is legitimate in other contexts. For example, anecdotal evidence from a friend about how to treat a certain disease constitutes empirical evidence that this treatment works but would not be considered scientific evidence. Others have argued that the traditional empiricist definition of empirical evidence as perceptual evidence is too narrow for much of scientific practice, which uses evidence from various kinds of non-perceptual equipment.

    Central to scientific evidence is that it was arrived at by following scientific method in the context of some scientific theory. But people rely on various forms of empirical evidence in their everyday lives that have not been obtained this way and therefore do not qualify as scientific evidence. One problem with non-scientific evidence is that it is less reliable, for example, due to cognitive biases like the anchoring effect, in which information obtained earlier is given more weight, although science done poorly is also subject to such biases, as in the example of p-hacking.

    Observation, experimentation and scientific method

    In the philosophy of science, it is sometimes held that there are two sources of empirical evidence: observation and experimentation. The idea behind this distinction is that only experimentation involves manipulation or intervention: phenomena are actively created instead of being passively observed. For example, inserting viral DNA into a bacterium is a form of experimentation while studying planetary orbits through a telescope belongs to mere observation. In these cases, the mutated DNA was actively produced by the biologist while the planetary orbits are independent of the astronomer observing them. Applied to the history of science, it is sometimes held that ancient science is mainly observational while the emphasis on experimentation is only present in modern science and responsible for the scientific revolution. This is sometimes phrased through the expression that modern science actively “puts questions to nature”. This distinction also underlies the categorization of sciences into experimental sciences, like physics, and observational sciences, like astronomy. While the distinction is relatively intuitive in paradigmatic cases, it has proven difficult to give a general definition of “intervention” applying to all cases, which is why it is sometimes outright rejected.

    Empirical evidence is required for a hypothesis to gain acceptance in the scientific community. Normally, this validation is achieved by the scientific method of forming a hypothesis, experimental design, peer review, reproduction of results, conference presentation, and journal publication. This requires rigorous communication of hypothesis (usually expressed in mathematics), experimental constraints and controls (expressed in terms of standard experimental apparatus), and a common understanding of measurement. In the scientific context, the term semi-empirical is used for qualifying theoretical methods that use, in part, basic axioms or postulated scientific laws and experimental results. Such methods are opposed to theoretical ab initio methods, which are purely deductive and based on first principles. Typical examples of both ab initio and semi-empirical methods can be found in computational chemistry.

  • MAMIL Meaning

    Mamil (or MAMIL) is an acronym and a pejorative term for a “middle-aged man in Lycra” – that is, men who ride an expensive racing bicycle for leisure, while wearing body-hugging jerseys and bicycle shorts.

    The word was reportedly coined by British marketing research firm Mintel in 2010. It gained further popularity in the United Kingdom with the success of Bradley Wiggins in the 2012 Tour de France and at the 2012 Summer Olympics, held in London. The British UCI World Championships victories in recent years have also spurred interest in cycling in the UK.

    In Australia the popularity of this sort of cycling has been associated with the Tour Down Under and the 2011 Tour de France winner Cadel Evans. Former Prime Minister Tony Abbott has been described as a “mamil”

    And in Slovakia, for example, the popularity of racing cycling and wearing colorful Lycra on the roads rose after Peter Sagan begun winning in Tour de France and World championships.

    Buying an expensive road bicycle has been described as a more healthy and affordable response to a midlife crisis than buying an expensive sports car.

    There are documentaries investigating this cycling culture. MAMIL is the title of a one-man play by New Zealand playwright Greg Cooper, written for actor Mark Hadlow. It is also the title of a feature-length documentary directed by Nickolas Bird and produced by Bird, Eleanor Sharpe and Mark Bird.

    Thirty or 40 years ago, people would ride a bike for economic reasons, but our research suggests that nowadays a bicycle is more a lifestyle addition, a way of demonstrating how affluent you are.

    Spring will usually bring Mamils out of hibernation in search of fresh Lycra and carbon fiber – their staple diet.

  • Scientific method

    The scientific method is an empirical method for acquiring knowledge that has been referred to while doing science since at least the 17th century. Historically, it was developed through the centuries from the ancient and medieval world. The scientific method involves careful observation coupled with rigorous skepticism, because cognitive assumptions can distort the interpretation of the observation. Scientific inquiry includes creating a testable hypothesis through inductive reasoning, testing it through experiments and statistical analysis, and adjusting or discarding the hypothesis based on the results.

    Although procedures vary across fields, the underlying process is often similar. In more detail: the scientific method involves making conjectures (hypothetical explanations), predicting the logical consequences of hypothesis, then carrying out experiments or empirical observations based on those predictions. A hypothesis is a conjecture based on knowledge obtained while seeking answers to the question. Hypotheses can be very specific or broad but must be falsifiable, implying that it is possible to identify a possible outcome of an experiment or observation that conflicts with predictions deduced from the hypothesis; otherwise, the hypothesis cannot be meaningfully tested.

    While the scientific method is often presented as a fixed sequence of steps, it actually represents a set of general principles. Not all steps take place in every scientific inquiry (nor to the same degree), and they are not always in the same order.Numerous discoveries have not followed the textbook model of the scientific method and chance has played a role, for instance.

    History

    The history of the scientific method considers changes in the methodology of scientific inquiry, not the history of science itself. The development of rules for scientific reasoning has not been straightforward; the scientific method has been the subject of intense and recurring debate throughout the history of science, and eminent natural philosophers and scientists have argued for the primacy of various approaches to establishing scientific knowledge.

    Different early expressions of empiricism and the scientific method can be found throughout history, for instance with the ancient Stoics, Aristotle, Epicurus, Alhazen, Avicenna, Al-Biruni, Roger Bacon, and William of Ockham.

    In the scientific revolution of the 16th and 17th centuries, some of the most important developments were the furthering of empiricism by Francis Bacon and Robert Hooke, the rationalist approach described by René Descartes, and inductivism, brought to particular prominence by Isaac Newton and those who followed him. Experiments were advocated by Francis Bacon and performed by Giambattista della Porta, Johannes Kepler, and Galileo Galilei. There was particular development aided by theoretical works by the skeptic Francisco Sanches, by idealists as well as empiricists John Locke, George Berkeley, and David Hume. C. S. Peirce formulated the hypothetico-deductive model in the 20th century, and the model has undergone significant revision since.

    The term “scientific method” emerged in the 19th century, as a result of significant institutional development of science, and terminologies establishing clear boundaries between science and non-science, such as “scientist” and “pseudoscience”. Throughout the 1830s and 1850s, when Baconianism was popular, naturalists like William Whewell, John Herschel, and John Stuart Mill engaged in debates over “induction” and “facts,” and were focused on how to generate knowledge. In the late 19th and early 20th centuries, a debate over realism vs. antirealism was conducted as powerful scientific theories extended beyond the realm of the observable.

    Modern use and critical thought

    The term “scientific method” came into popular use in the twentieth century; Dewey’s 1910 book, How We Think, inspired popular guidelines.It appeared in dictionaries and science textbooks, although there was little consensus on its meaning. Although there was growth through the middle of the twentieth century, by the 1960s and 1970s numerous influential philosophers of science such as Thomas Kuhn and Paul Feyerabend had questioned the universality of the “scientific method,” and largely replaced the notion of science as a homogeneous and universal method with that of it being a heterogeneous and local practice. In particular, Paul Feyerabend, in the 1975 first edition of his book Against Method, argued against there being any universal rules of science; Karl Popper, and Gauch 2003,[6] disagreed with Feyerabend’s claim.

    Later stances include physicist Lee Smolin’s 2013 essay “There Is No Scientific Method”, in which he espouses two ethical principles, and historian of science Daniel Thurs’ chapter in the 2015 book Newton’s Apple and Other Myths about Science, which concluded that the scientific method is a myth or, at best, an idealization. As myths are beliefs, they are subject to the narrative fallacy, as pointed out by Taleb. Philosophers Robert Nola and Howard Sankey, in their 2007 book Theories of Scientific Method, said that debates over the scientific method continue, and argued that Feyerabend, despite the title of Against Method, accepted certain rules of method and attempted to justify those rules with a meta methodology. Staddon (2017) argues it is a mistake to try following rules in the absence of an algorithmic scientific method; in that case, “science is best understood through examples”.But algorithmic methods, such as disproof of existing theory by experiment have been used since Alhacen (1027) and his Book of Optics and Galileo (1638) and his Two New Sciences, and The Assayer, which still stand as scientific method.

    Elements of inquiry

    Overview

    The scientific method is the process by which science is carried out. As in other areas of inquiry, science (through the scientific method) can build on previous knowledge, and unify understanding of its studied topics over time. This model can be seen to underlie the scientific revolution.

    The overall process involves making conjectures (hypotheses), predicting their logical consequences, then carrying out experiments based on those predictions to determine whether the original conjecture was correct. However, there are difficulties in a formulaic statement of method. Though the scientific method is often presented as a fixed sequence of steps, these actions are more accurately general principles. Not all steps take place in every scientific inquiry (nor to the same degree), and they are not always done in the same order.

    Factors of scientific inquiry

    There are different ways of outlining the basic method used for scientific inquiry. The scientific community and philosophers of science generally agree on the following classification of method components. These methodological elements and organization of procedures tend to be more characteristic of experimental sciences than social sciences. Nonetheless, the cycle of formulating hypotheses, testing and analyzing the results, and formulating new hypotheses, will resemble the cycle described below.The scientific method is an iterative, cyclical process through which information is continually revised. It is generally recognized to develop advances in knowledge through the following elements, in varying combinations or contributions :

    Characterizations (observations, definitions, and measurements of the subject of inquiry)
    Hypotheses (theoretical, hypothetical explanations of observations and measurements of the subject)
    Predictions (inductive and deductive reasoning from the hypothesis or theory)
    Experiments (tests of all of the above)
    Each element of the scientific method is subject to peer review for possible mistakes. These activities do not describe all that scientists do but apply mostly to experimental sciences (e.g., physics, chemistry, biology, and psychology). The elements above are often taught in the educational system as “the scientific method”.

    The scientific method is not a single recipe: it requires intelligence, imagination, and creativity. In this sense, it is not a mindless set of standards and procedures to follow but is rather an ongoing cycle, constantly developing more useful, accurate, and comprehensive models and methods. For example, when Einstein developed the Special and General Theories of Relativity, he did not in any way refute or discount Newton’s Principia. On the contrary, if the astronomically massive, the feather-light, and the extremely fast are removed from Einstein’s theories – all phenomena Newton could not have observed – Newton’s equations are what remain. Einstein’s theories are expansions and refinements of Newton’s theories and, thus, increase confidence in Newton’s work.

    An iterative, pragmatic scheme of the four points above is sometimes offered as a guideline for proceeding:

    Define a question
    Gather information and resources (observe)
    Form an explanatory hypothesis
    Test the hypothesis by performing an experiment and collecting data in a reproducible manner
    Analyze the data
    Interpret the data and draw conclusions that serve as a starting point for a new hypothesis
    Publish results
    Retest (frequently done by other scientists)
    The iterative cycle inherent in this step-by-step method goes from point 3 to 6 and back to 3 again.

    While this schema outlines a typical hypothesis/testing method, many philosophers, historians, and sociologists of science, including Paul Feyerabend, claim that such descriptions of scientific method have little relation to the ways that science is actually practiced.

    Characterizations

    The basic elements of the scientific method are illustrated by the following example (which occurred from 1944 to 1953) from the discovery of the structure of DNA (marked with DNA label and indented).

    DNA label In 1950, it was known that genetic inheritance had a mathematical description, starting with the studies of Gregor Mendel, and that DNA contained genetic information (Oswald Avery’s transforming principle). But the mechanism of storing genetic information (i.e., genes) in DNA was unclear. Researchers in Bragg’s laboratory at Cambridge University made X-ray diffraction pictures of various molecules, starting with crystals of salt, and proceeding to more complicated substances. Using clues painstakingly assembled over decades, beginning with its chemical composition, it was determined that it should be possible to characterize the physical structure of DNA, and the X-ray images would be the vehicle.

    The scientific method depends upon increasingly sophisticated characterizations of the subjects of investigation. (The subjects can also be called unsolved problems or the unknowns.) For example, Benjamin Franklin conjectured, correctly, that St. Elmo’s fire was electrical in nature, but it has taken a long series of experiments and theoretical changes to establish this. While seeking the pertinent properties of the subjects, careful thought may also entail some definitions and observations; these observations often demand careful measurements and/or counting can take the form of expansive empirical research.

    A scientific question can refer to the explanation of a specific observation, as in “Why is the sky blue?” but can also be open-ended, as in “How can I design a drug to cure this particular disease?” This stage frequently involves finding and evaluating evidence from previous experiments, personal scientific observations or assertions, as well as the work of other scientists. If the answer is already known, a different question that builds on the evidence can be posed. When applying the scientific method to research, determining a good question can be very difficult and it will affect the outcome of the investigation.

    The systematic, careful collection of measurements or counts of relevant quantities is often the critical difference between pseudo-sciences, such as alchemy, and science, such as chemistry or biology. Scientific measurements are usually tabulated, graphed, or mapped, and statistical manipulations, such as correlation and regression, performed on them. The measurements might be made in a controlled setting, such as a laboratory, or made on more or less inaccessible or unmanipulatable objects such as stars or human populations. The measurements often require specialized scientific instruments such as thermometers, spectroscopes, particle accelerators, or voltmeters, and the progress of a scientific field is usually intimately tied to their invention and improvement.

    I am not accustomed to saying anything with certainty after only one or two observations.

    — Andreas Vesalius (1546)

    Definition

    The scientific definition of a term sometimes differs substantially from its natural language usage. For example, mass and weight overlap in meaning in common discourse, but have distinct meanings in mechanics. Scientific quantities are often characterized by their units of measure which can later be described in terms of conventional physical units when communicating the work.

    New theories are sometimes developed after realizing certain terms have not previously been sufficiently clearly defined. For example, Albert Einstein’s first paper on relativity begins by defining simultaneity and the means for determining length. These ideas were skipped over by Isaac Newton with, “I do not define time, space, place and motion, as being well known to all.” Einstein’s paper then demonstrates that they (viz., absolute time and length independent of motion) were approximations. Francis Crick cautions us that when characterizing a subject, however, it can be premature to define something when it remains ill-understood. In Crick’s study of consciousness, he actually found it easier to study awareness in the visual system, rather than to study free will, for example. His cautionary example was the gene; the gene was much more poorly understood before Watson and Crick’s pioneering discovery of the structure of DNA; it would have been counterproductive to spend much time on the definition of the gene, before them.

    Hypothesis development

    DNA label Linus Pauling proposed that DNA might be a triple helix. This hypothesis was also considered by Francis Crick and James D. Watson but discarded. When Watson and Crick learned of Pauling’s hypothesis, they understood from existing data that Pauling was wrong. and that Pauling would soon admit his difficulties with that structure.

    A hypothesis is a suggested explanation of a phenomenon, or alternately a reasoned proposal suggesting a possible correlation between or among a set of phenomena. Normally, hypotheses have the form of a mathematical model. Sometimes, but not always, they can also be formulated as existential statements, stating that some particular instance of the phenomenon being studied has some characteristic and causal explanations, which have the general form of universal statements, stating that every instance of the phenomenon has a particular characteristic.

    Scientists are free to use whatever resources they have – their own creativity, ideas from other fields, inductive reasoning, Bayesian inference, and so on – to imagine possible explanations for a phenomenon under study. Albert Einstein once observed that “there is no logical bridge between phenomena and their theoretical principles.” Charles Sanders Peirce, borrowing a page from Aristotle (Prior Analytics, 2.25) described the incipient stages of inquiry, instigated by the “irritation of doubt” to venture a plausible guess, as abductive reasoning.II,p.290 The history of science is filled with stories of scientists claiming a “flash of inspiration”, or a hunch, which then motivated them to look for evidence to support or refute their idea. Michael Polanyi made such creativity the centerpiece of his discussion of methodology.

    William Glen observes that

    the success of a hypothesis, or its service to science, lies not simply in its perceived “truth”, or power to displace, subsume or reduce a predecessor idea, but perhaps more in its ability to stimulate the research that will illuminate … bald suppositions and areas of vagueness.

    — William Glen, The Mass-Extinction Debates


    In general, scientists tend to look for theories that are “elegant” or “beautiful”. Scientists often use these terms to refer to a theory that is following the known facts but is nevertheless relatively simple and easy to handle. Occam’s Razor serves as a rule of thumb for choosing the most desirable amongst a group of equally explanatory hypotheses.

    To minimize the confirmation bias that results from entertaining a single hypothesis, strong inference emphasizes the need for entertaining multiple alternative hypotheses, and avoiding artifacts.

    Predictions from the hypothesis

    DNA label James D. Watson, Francis Crick, and others hypothesized that DNA had a helical structure. This implied that DNA’s X-ray diffraction pattern would be ‘x shaped’. This prediction followed from the work of Cochran, Crick and Vand (and independently by Stokes). The Cochran-Crick-Vand-Stokes theorem provided a mathematical explanation for the empirical observation that diffraction from helical structures produces x-shaped patterns. In their first paper, Watson and Crick also noted that the double helix structure they proposed provided a simple mechanism for DNA replication, writing, “It has not escaped our notice that the specific pairing we have postulated immediately suggests a possible copying mechanism for the genetic material”.

    Any useful hypothesis will enable predictions, by reasoning including deductive reasoning. It might predict the outcome of an experiment in a laboratory setting or the observation of a phenomenon in nature. The prediction can also be statistical and deal only with probabilities.

    It is essential that the outcome of testing such a prediction be currently unknown. Only in this case does a successful outcome increase the probability that the hypothesis is true. If the outcome is already known, it is called a consequence and should have already been considered while formulating the hypothesis.

    If the predictions are not accessible by observation or experience, the hypothesis is not yet testable and so will remain to that extent unscientific in a strict sense. A new technology or theory might make the necessary experiments feasible. For example, while a hypothesis on the existence of other intelligent species may be convincing with scientifically based speculation, no known experiment can test this hypothesis. Therefore, science itself can have little to say about the possibility. In the future, a new technique may allow for an experimental test and the speculation would then become part of accepted science.

    For example, Einstein’s theory of general relativity makes several specific predictions about the observable structure of spacetime, such as that light bends in a gravitational field, and that the amount of bending depends in a precise way on the strength of that gravitational field. Arthur Eddington’s observations made during a 1919 solar eclipse supported General Relativity rather than Newtonian gravitation.

    Experiments

    DNA label Watson and Crick showed an initial (and incorrect) proposal for the structure of DNA to a team from King’s College London – Rosalind Franklin, Maurice Wilkins, and Raymond Gosling. Franklin immediately spotted the flaws which concerned the water content. Later Watson saw Franklin’s photo 51, a detailed X-ray diffraction image, which showed an X-shape and was able to confirm the structure was helical.

    Once predictions are made, they can be sought by experiments. If the test results contradict the predictions, the hypotheses which entailed them are called into question and become less tenable. Sometimes the experiments are conducted incorrectly or are not very well designed when compared to a crucial experiment. If the experimental results confirm the predictions, then the hypotheses are considered more likely to be correct, but might still be wrong and continue to be subject to further testing. The experimental control is a technique for dealing with observational error. This technique uses the contrast between multiple samples, or observations, or populations, under differing conditions, to see what varies or what remains the same. We vary the conditions for the acts of measurement, to help isolate what has changed. Mill’s canons can then help us figure out what the important factor is. Factor analysis is one technique for discovering the important factor in an effect.

    Depending on the predictions, the experiments can have different shapes. It could be a classical experiment in a laboratory setting, a double-blind study or an archaeological excavation. Even taking a plane from New York to Paris is an experiment that tests the aerodynamical hypotheses used for constructing the plane.

    These institutions thereby reduce the research function to a cost/benefit, which is expressed as money, and the time and attention of the researchers to be expended, in exchange for a report to their constituents. Current large instruments, such as CERN’s Large Hadron Collider (LHC), or LIGO, or the National Ignition Facility (NIF), or the International Space Station (ISS), or the James Webb Space Telescope (JWST), entail expected costs of billions of dollars, and timeframes extending over decades. These kinds of institutions affect public policy, on a national or even international basis, and the researchers would require shared access to such machines and their adjunct infrastructure.

    Scientists assume an attitude of openness and accountability on the part of those experimenting. Detailed record-keeping is essential, to aid in recording and reporting on the experimental results, and supports the effectiveness and integrity of the procedure. They will also assist in reproducing the experimental results, likely by others. Traces of this approach can be seen in the work of Hipparchus (190–120 BCE), when determining a value for the precession of the Earth, while controlled experiments can be seen in the works of al-Battani (853–929 CE) and Alhazen (965–1039 CE).

    Communication and iteration

    DNA label Watson and Crick then produced their model, using this information along with the previously known information about DNA’s composition, especially Chargaff’s rules of base pairing. After considerable fruitless experimentation, being discouraged by their superior from continuing, and numerous false starts,Watson and Crick were able to infer the essential structure of DNA by concrete modeling of the physical shapes of the nucleotides which comprise it.They were guided by the bond lengths which had been deduced by Linus Pauling and by Rosalind Franklin’s X-ray diffraction images.

    The scientific method is iterative. At any stage, it is possible to refine its accuracy and precision, so that some consideration will lead the scientist to repeat an earlier part of the process. Failure to develop an interesting hypothesis may lead a scientist to re-define the subject under consideration. Failure of a hypothesis to produce interesting and testable predictions may lead to reconsideration of the hypothesis or of the definition of the subject. Failure of an experiment to produce interesting results may lead a scientist to reconsider the experimental method, the hypothesis, or the definition of the subject.

    This manner of iteration can span decades and sometimes centuries. Published papers can be built upon. For example: By 1027, Alhazen, based on his measurements of the refraction of light, was able to deduce that outer space was less dense than air, that is: “the body of the heavens is rarer than the body of air”. In 1079 Ibn Mu’adh’s Treatise On Twilight was able to infer that Earth’s atmosphere was 50 miles thick, based on atmospheric refraction of the sun’s rays.

    This is why the scientific method is often represented as circular – new information leads to new characterisations, and the cycle of science continues. Measurements collected can be archived, passed onwards and used by others. Other scientists may start their own research and enter the process at any stage. They might adopt the characterization and formulate their own hypothesis, or they might adopt the hypothesis and deduce their own predictions. Often the experiment is not done by the person who made the prediction, and the characterization is based on experiments done by someone else. Published results of experiments can also serve as a hypothesis predicting their own reproducibility.

    Confirmation

    Science is a social enterprise, and scientific work tends to be accepted by the scientific community when it has been confirmed. Crucially, experimental and theoretical results must be reproduced by others within the scientific community. Researchers have given their lives for this vision; Georg Wilhelm Richmann was killed by ball lightning (1753) when attempting to replicate the 1752 kite-flying experiment of Benjamin Franklin.

    If an experiment cannot be repeated to produce the same results, this implies that the original results might have been in error. As a result, it is common for a single experiment to be performed multiple times, especially when there are uncontrolled variables or other indications of experimental error. For significant or surprising results, other scientists may also attempt to replicate the results for themselves, especially if those results would be important to their own work. Replication has become a contentious issue in social and biomedical science where treatments are administered to groups of individuals. Typically an experimental group gets the treatment, such as a drug, and the control group gets a placebo. John Ioannidis in 2005 pointed out that the method being used has led to many findings that cannot be replicated.

    The process of peer review involves the evaluation of the experiment by experts, who typically give their opinions anonymously. Some journals request that the experimenter provide lists of possible peer reviewers, especially if the field is highly specialized. Peer review does not certify the correctness of the results, only that, in the opinion of the reviewer, the experiments themselves were sound (based on the description supplied by the experimenter). If the work passes peer review, which occasionally may require new experiments requested by the reviewers, it will be published in a peer-reviewed scientific journal. The specific journal that publishes the results indicates the perceived quality of the work.

    Scientists typically are careful in recording their data, a requirement promoted by Ludwik Fleck (1896–1961) and others. Though not typically required, they might be requested to supply this data to other scientists who wish to replicate their original results (or parts of their original results), extending to the sharing of any experimental samples that may be difficult to obtain. To protect against bad science and fraudulent data, government research-granting agencies such as the National Science Foundation, and science journals, including Nature and Science, have a policy that researchers must archive their data and methods so that other researchers can test the data and methods and build on the research that has gone before. Scientific data archiving can be done at several national archives in the U.S. or the World Data Center.

    Foundational principles

    Honesty, openness, and falsifiability

    The unfettered principles of science are to strive for accuracy and the creed of honesty; openness already being a matter of degrees. Openness is restricted by the general rigour of scepticism. And of course the matter of non-science.

    Smolin, in 2013, espoused ethical principles rather than giving any potentially limited definition of the rules of inquiry. His ideas stand in the context of the scale of data–driven and big science, which has seen increased importance of honesty and consequently reproducibility. His thought is that science is a community effort by those who have accreditation and are working within the community. He also warns against overzealous parsimony.

    Popper previously took ethical principles even further, going as far as to ascribe value to theories only if they were falsifiable. Popper used the falsifiability criterion to demarcate a scientific theory from a theory like astrology: both “explain” observations, but the scientific theory takes the risk of making predictions that decide whether it is right or wrong:

    “Those among us who are unwilling to expose their ideas to the hazard of refutation do not take part in the game of science.”

    — Karl Popper, The Logic of Scientific Discovery (2002 [1935])

    Theory’s interactions with observation

    Science has limits. Those limits are usually deemed to be answers to questions that aren’t in science’s domain, such as faith. Science has other limits as well, as it seeks to make true statements about reality. The nature of truth and the discussion on how scientific statements relate to reality is best left to the article on the philosophy of science here. More immediately topical limitations show themselves in the observation of reality.

    It is the natural limitations of scientific inquiry that there is no pure observation as theory is required to interpret empirical data, and observation is therefore influenced by the observer’s conceptual framework. As science is an unfinished project, this does lead to difficulties. Namely, that false conclusions are drawn, because of limited information.

    An example here are the experiments of Kepler and Brahe, used by Hanson to illustrate the concept. Despite observing the same sunrise the two scientists came to different conclusions—their intersubjectivity leading to differing conclusions. Johannes Kepler used Tycho Brahe’s method of observation, which was to project the image of the Sun on a piece of paper through a pinhole aperture, instead of looking directly at the Sun. He disagreed with Brahe’s conclusion that total eclipses of the Sun were impossible because, contrary to Brahe, he knew that there were historical accounts of total eclipses. Instead, he deduced that the images taken would become more accurate, the larger the aperture—this fact is now fundamental for optical system design. Another historic example here is the discovery of Neptune, credited as being found via mathematics because previous observers didn’t know what they were looking at.

    Empiricism, rationalism, and more pragmatic views

    Scientific endeavour can be characterised as the pursuit of truths about the natural world or as the elimination of doubt about the same. The former is the direct construction of explanations from empirical data and logic, the latter the reduction of potential explanations. It was established above how the interpretation of empirical data is theory-laden, so neither approach is trivial.

    The ubiquitous element in the scientific method is empiricism, which holds that knowledge is created by a process involving observation; scientific theories generalize observations. This is in opposition to stringent forms of rationalism, which holds that knowledge is created by the human intellect; later clarified by Popper to be built on prior theory. The scientific method embodies the position that reason alone cannot solve a particular scientific problem; it unequivocally refutes claims that revelation, political or religious dogma, appeals to tradition, commonly held beliefs, common sense, or currently held theories pose the only possible means of demonstrating truth.

    In 1877, C. S. Peirce characterized inquiry in general not as the pursuit of truth per se but as the struggle to move from irritating, inhibitory doubts born of surprises, disagreements, and the like, and to reach a secure belief, the belief being that on which one is prepared to act. His pragmatic views framed scientific inquiry as part of a broader spectrum and as spurred, like inquiry generally, by actual doubt, not mere verbal or “hyperbolic doubt”, which he held to be fruitless. This “hyperbolic doubt” Peirce argues against here is of course just another name for Cartesian doubt associated with René Descartes. It is a methodological route to certain knowledge by identifying what can’t be doubted.

    A strong formulation of the scientific method is not always aligned with a form of empiricism in which the empirical data is put forward in the form of experience or other abstracted forms of knowledge as in current scientific practice the use of scientific modelling and reliance on abstract typologies and theories is normally accepted. In 2010, Hawking suggested that physics’ models of reality should simply be accepted where they prove to make useful predictions. He calls the concept model-dependent realism.

    Rationality

    Rationality embodies the essence of sound reasoning, a cornerstone not only in philosophical discourse but also in the realms of science and practical decision-making. According to the traditional viewpoint, rationality serves a dual purpose: it governs beliefs, ensuring they align with logical principles, and it steers actions, directing them towards coherent and beneficial outcomes. This understanding underscores the pivotal role of reason in shaping our understanding of the world and in informing our choices and behaviours. The following section will first explore beliefs and biases, and then get to the rational reasoning most associated with the sciences.

    Beliefs and biases

    Scientific methodology often directs that hypotheses be tested in controlled conditions wherever possible. This is frequently possible in certain areas, such as in the biological sciences, and more difficult in other areas, such as in astronomy.

    The practice of experimental control and reproducibility can have the effect of diminishing the potentially harmful effects of circumstance, and to a degree, personal bias. For example, pre-existing beliefs can alter the interpretation of results, as in confirmation bias; this is a heuristic that leads a person with a particular belief to see things as reinforcing their belief, even if another observer might disagree (in other words, people tend to observe what they expect to observe).

    The action of thought is excited by the irritation of doubt, and ceases when belief is attained.

    — C.S. Peirce, How to Make Our Ideas Clear (1877)


    A historical example is the belief that the legs of a galloping horse are splayed at the point when none of the horse’s legs touch the ground, to the point of this image being included in paintings by its supporters. However, the first stop-action pictures of a horse’s gallop by Eadweard Muybridge showed this to be false, and that the legs are instead gathered together.

    Another important human bias that plays a role is a preference for new, surprising statements (see Appeal to novelty), which can result in a search for evidence that the new is true. Poorly attested beliefs can be believed and acted upon via a less rigorous heuristic.

    Goldhaber and Nieto published in 2010 the observation that if theoretical structures with “many closely neighboring subjects are described by connecting theoretical concepts, then the theoretical structure acquires a robustness which makes it increasingly hard – though certainly never impossible – to overturn”. When a narrative is constructed its elements become easier to believe.

    Fleck (1979), p. 27 notes “Words and ideas are originally phonetic and mental equivalences of the experiences coinciding with them. … Such proto-ideas are at first always too broad and insufficiently specialized. … Once a structurally complete and closed system of opinions consisting of many details and relations has been formed, it offers enduring resistance to anything that contradicts it”. Sometimes, these relations have their elements assumed a priori, or contain some other logical or methodological flaw in the process that ultimately produced them. Donald M. MacKay has analyzed these elements in terms of limits to the accuracy of measurement and has related them to instrumental elements in a category of measurement.

    Deductive and inductive reasoning

    The idea of there being two opposed justifications for truth has shown up throughout the history of scientific method as analysis versus synthesis, non-ampliative/ampliative, or even confirmation and verification. (And there are other kinds of reasoning.) One to use what is observed to build towards fundamental truths – and the other to derive from those fundamental truths more specific principles.

    Deductive reasoning is the building of knowledge based on what has been shown to be true before. It requires the assumption of fact established prior, and, given the truth of the assumptions, a valid deduction guarantees the truth of the conclusion. Inductive reasoning builds knowledge not from established truth, but from a body of observations. It requires stringent scepticism regarding observed phenomena, because cognitive assumptions can distort the interpretation of initial perceptions.

    An example for how inductive and deductive reasoning works can be found in the history of gravitational theory. It took thousands of years of measurements, from the Chaldean, Indian, Persian, Greek, Arabic, and European astronomers, to fully record the motion of planet Earth.Kepler(and others) were then able to build their early theories by generalizing the collected data inductively, and Newton was able to unify prior theory and measurements into the consequences of his laws of motion in 1727.

    Another common example of inductive reasoning is the observation of a counterexample to current theory inducing the need for new ideas. Le Verrier in 1859 pointed out problems with the perihelion of Mercury that showed Newton’s theory to be at least incomplete. The observed difference of Mercury’s precession between Newtonian theory and observation was one of the things that occurred to Einstein as a possible early test of his theory of relativity. His relativistic calculations matched observation much more closely than Newtonian theory did. Though, today’s Standard Model of physics suggests that we still do not know at least some of the concepts surrounding Einstein’s theory, it holds to this day and is being built on deductively.

    A theory being assumed as true and subsequently built on is a common example of deductive reasoning. Theory building on Einstein’s achievement can simply state that ‘we have shown that this case fulfils the conditions under which general/special relativity applies, therefore its conclusions apply also’. If it was properly shown that ‘this case’ fulfils the conditions, the conclusion follows. An extension of this is the assumption of a solution to an open problem. This weaker kind of deductive reasoning will get used in current research, when multiple scientists or even teams of researchers are all gradually solving specific cases in working towards proving a larger theory. This often sees hypotheses being revised again and again as new proof emerges.

    This way of presenting inductive and deductive reasoning shows part of why science is often presented as being a cycle of iteration. It is important to keep in mind that that cycle’s foundations lie in reasoning, and not wholly in the following of procedure.

    Certainty, probabilities, and statistical inference

    Claims of scientific truth can be opposed in three ways: by falsifying them, by questioning their certainty, or by asserting the claim itself to be incoherent. Incoherence, here, means internal errors in logic, like stating opposites to be true; falsification is what Popper would have called the honest work of conjecture and refutation — certainty, perhaps, is where difficulties in telling truths from non-truths arise most easily.

    Measurements in scientific work are usually accompanied by estimates of their uncertainty. The uncertainty is often estimated by making repeated measurements of the desired quantity. Uncertainties may also be calculated by consideration of the uncertainties of the individual underlying quantities used. Counts of things, such as the number of people in a nation at a particular time, may also have an uncertainty due to data collection limitations. Or counts may represent a sample of desired quantities, with an uncertainty that depends upon the sampling method used and the number of samples taken.

    In the case of measurement imprecision, there will simply be a ‘probable deviation’ expressing itself in a study’s conclusions. Statistics are different. Inductive statistical generalisation will take sample data and extrapolate more general conclusions, which has to be justified — and scrutinised. It can even be said that statistical models are only ever useful, but never a complete representation of circumstances.

    In statistical analysis, expected and unexpected bias is a large factor. Research questions, the collection of data, or the interpretation of results, all are subject to larger amounts of scrutiny than in comfortably logical environments. Statistical models go through a process for validation, for which one could even say that awareness of potential biases is more important than the hard logic; errors in logic are easier to find in peer review, after all.[u] More general, claims to rational knowledge, and especially statistics, have to be put into their appropriate context. Simple statements such as ‘9 out of 10 doctors recommend’ are therefore of unknown quality because they do not justify their methodology.

    Lack of familiarity with statistical methodologies can result in erroneous conclusions. Foregoing the easy example, multiple probabilities interacting is where, for example medical professionals, have shown a lack of proper understanding. Bayes’ theorem is the mathematical principle lining out how standing probabilities are adjusted given new information. The boy or girl paradox is a common example. In knowledge representation, Bayesian estimation of mutual information between random variables is a way to measure dependence, independence, or interdependence of the information under scrutiny.

    Beyond commonly associated survey methodology of field research, the concept together with probabilistic reasoning is used to advance fields of science where research objects have no definitive states of being. For example, in statistical mechanics.

    Methods of inquiry

    Hypothetico-deductive method

    The hypothetico-deductive model, or hypothesis-testing method, or “traditional” scientific method is, as the name implies, based on the formation of hypotheses and their testing via deductive reasoning. A hypothesis stating implications, often called predictions, that are falsifiable via experiment is of central importance here, as not the hypothesis but its implications are what is tested. Basically, scientists will look at the hypothetical consequences a (potential) theory holds and prove or disprove those instead of the theory itself. If an experimental test of those hypothetical consequences shows them to be false, it follows logically that the part of the theory that implied them was false also. If they show as true however, it does not prove the theory definitively.

    The logic of this testing is what affords this method of inquiry to be reasoned deductively. The formulated hypothesis is assumed to be ‘true’, and from that ‘true’ statement implications are inferred. If the following tests show the implications to be false, it follows that the hypothesis was false also. If test show the implications to be true, new insights will be gained. It is important to be aware that a positive test here will at best strongly imply but not definitively prove the tested hypothesis, as deductive inference (A ⇒ B) is not equivalent like that; only (¬B ⇒ ¬A) is valid logic. Their positive outcomes however, as Hempel put it, provide “at least some support, some corroboration or confirmation for it”. This is why Popper insisted on fielded hypotheses to be falsifieable, as successful tests imply very little otherwise. As Gillies put it, “successful theories are those that survive elimination through falsification”.

    Deductive reasoning in this mode of inquiry will sometimes be replaced by abductive reasoning—the search for the most plausible explanation via logical inference. For example, in biology, where general laws are few, as valid deductions rely on solid presuppositions.

    Inductive method

    The inductivist approach to deriving scientific truth first rose to prominence with Francis Bacon and particularly with Isaac Newton and those who followed him. After the establishment of the HD-method, it was often put aside as something of a “fishing expedition” though. It is still valid to some degree, but today’s inductive method is often far removed from the historic approach—the scale of the data collected lending new effectiveness to the method. It is most-associated with data-mining projects or large-scale observation projects. In both these cases, it is often not at all clear what the results of proposed experiments will be, and thus knowledge will arise after the collection of data through inductive reasoning.

    Where the traditional method of inquiry does both, the inductive approach usually formulates only a research question, not a hypothesis. Following the initial question instead, a suitable “high-throughput method” of data-collection is determined, the resulting data processed and ‘cleaned up’, and conclusions drawn after. “This shift in focus elevates the data to the supreme role of revealing novel insights by themselves”.

    The advantage the inductive method has over methods formulating a hypothesis that it is essentially free of “a researcher’s preconceived notions” regarding their subject. On the other hand, inductive reasoning is always attached to a measure of certainty, as all inductively reasoned conclusions are. This measure of certainty can reach quite high degrees, though. For example, in the determination of large primes, which are used in encryption software.

    Mathematical modelling

    Mathematical modelling, or allochthonous reasoning, typically is the formulation of a hypothesis followed by building mathematical constructs that can be tested in place of conducting physical laboratory experiments. This approach has two main factors: simplification/abstraction and secondly a set of correspondence rules. The correspondence rules lay out how the constructed model will relate back to reality-how truth is derived; and the simplifying steps taken in the abstraction of the given system are to reduce factors that do not bear relevance and thereby reduce unexpected errors. These steps can also help the researcher in understanding the important factors of the system, how far parsimony can be taken until the system becomes more and more unchangeable and thereby stable. Parsimony and related principles are further explored below.

    Once this translation into mathematics is complete, the resulting model, in place of the corresponding system, can be analysed through purely mathematical and computational means. The results of this analysis are of course also purely mathematical in nature and get translated back to the system as it exists in reality via the previously determined correspondence rules—iteration following review and interpretation of the findings. The way such models are reasoned will often be mathematically deductive—but they don’t have to be. An example here are Monte-Carlo simulations. These generate empirical data “arbitrarily”, and, while they may not be able to reveal universal principles, they can nevertheless be useful.

    Scientific inquiry

    Scientific inquiry generally aims to obtain knowledge in the form of testable explanations that scientists can use to predict the results of future experiments. This allows scientists to gain a better understanding of the topic under study, and later to use that understanding to intervene in its causal mechanisms (such as to cure disease). The better an explanation is at making predictions, the more useful it frequently can be, and the more likely it will continue to explain a body of evidence better than its alternatives. The most successful explanations – those that explain and make accurate predictions in a wide range of circumstances – are often called scientific theories.

    Most experimental results do not produce large changes in human understanding; improvements in theoretical scientific understanding typically result from a gradual process of development over time, sometimes across different domains of science. Scientific models vary in the extent to which they have been experimentally tested and for how long, and in their acceptance in the scientific community. In general, explanations become accepted over time as evidence accumulates on a given topic, and the explanation in question proves more powerful than its alternatives at explaining the evidence. Often subsequent researchers re-formulate the explanations over time, or combined explanations to produce new explanations.

    Properties of scientific inquiry

    Scientific knowledge is closely tied to empirical findings and can remain subject to falsification if new experimental observations are incompatible with what is found. That is, no theory can ever be considered final since new problematic evidence might be discovered. If such evidence is found, a new theory may be proposed, or (more commonly) it is found that modifications to the previous theory are sufficient to explain the new evidence. The strength of a theory relates to how long it has persisted without major alteration to its core principles.

    Theories can also become subsumed by other theories. For example, Newton’s laws explained thousands of years of scientific observations of the planets almost perfectly. However, these laws were then determined to be special cases of a more general theory (relativity), which explained both the (previously unexplained) exceptions to Newton’s laws and predicted and explained other observations such as the deflection of light by gravity. Thus, in certain cases independent, unconnected, scientific observations can be connected, unified by principles of increasing explanatory power.

    Since new theories might be more comprehensive than what preceded them, and thus be able to explain more than previous ones, successor theories might be able to meet a higher standard by explaining a larger body of observations than their predecessors. For example, the theory of evolution explains the diversity of life on Earth, how species adapt to their environments, and many other patterns observed in the natural world; its most recent major modification was unification with genetics to form the modern evolutionary synthesis. In subsequent modifications, it has also subsumed aspects of many other fields such as biochemistry and molecular biology.

    Heuristics

    Confirmation theory

    During the course of history, one theory has succeeded another, and some have suggested further work while others have seemed content just to explain the phenomena. The reasons why one theory has replaced another are not always obvious or simple. The philosophy of science includes the question: What criteria are satisfied by a ‘good’ theory. This question has a long history, and many scientists, as well as philosophers, have considered it. The objective is to be able to choose one theory as preferable to another without introducing cognitive bias. Though different thinkers emphasize different aspects, a good theory:

    is accurate (the trivial element);
    is consistent, both internally and with other relevant currently accepted theories;
    has explanatory power, meaning its consequences extend beyond the data it is required to explain;
    has unificatory power; as in its organizing otherwise confused and isolated phenomena
    and is fruitful for further research.
    In trying to look for such theories, scientists will, given a lack of guidance by empirical evidence, try to adhere to:

    parsimony in causal explanations
    and look for invariant observations.
    Scientists will sometimes also list the very subjective criteria of “formal elegance” which can indicate multiple different things.
    The goal here is to make the choice between theories less arbitrary. Nonetheless, these criteria contain subjective elements, and should be considered heuristics rather than a definitive. Also, criteria such as these do not necessarily decide between alternative theories. Quoting Bird:

    “[Such criteria] cannot determine scientific choice. First, which features of a theory satisfy these criteria may be disputable (e.g. does simplicity concern the ontological commitments of a theory or its mathematical form?). Secondly, these criteria are imprecise, and so there is room for disagreement about the degree to which they hold. Thirdly, there can be disagreement about how they are to be weighted relative to one another, especially when they conflict.”

    It also is debatable whether existing scientific theories satisfy all these criteria, which may represent goals not yet achieved. For example, explanatory power over all existing observations is satisfied by no one theory at the moment.

    Parsimony

    The desiderata of a “good” theory have been debated for centuries, going back perhaps even earlier than Occam’s razor, which is often taken as an attribute of a good theory. Science tries to be simple. When gathered data supports multiple explanations, the most simple explanation for phenomena or the most simple formation of a theory is recommended by the principle of parsimony. Scientists go as far as to call simple proofs of complex statements beautiful.

    We are to admit no more causes of natural things than such as are both true and sufficient to explain their appearances.

    — Isaac Newton, Philosophiæ Naturalis Principia Mathematica (1723 )
    The concept of parsimony should not be held to imply complete frugality in the pursuit of scientific truth. The general process starts at the opposite end of there being a vast number of potential explanations and general disorder. An example can be seen in Paul Krugman’s process, who makes explicit to “dare to be silly”. He writes that in his work on new theories of international trade he reviewed prior work with an open frame of mind and broadened his initial viewpoint even in unlikely directions. Once he had a sufficient body of ideas, he would try to simplify and thus find what worked among what did not. Specific to Krugman here was to “question the question”. He recognised that prior work had applied erroneous models to already present evidence, commenting that “intelligent commentary was ignored”. Thus touching on the need to bridge the common bias against other circles of thought.

    Elegance

    Occam’s razor might fall under the heading of “simple elegance”, but it is arguable that parsimony and elegance pull in different directions. Introducing additional elements could simplify theory formulation, whereas simplifying a theory’s ontology might lead to increased syntactical complexity.

    Sometimes ad-hoc modifications of a failing idea may also be dismissed as lacking “formal elegance”. This appeal to what may be called “aesthetic” is hard to characterise, but essentially about a sort of familiarity. Though, argument based on “elegance” is contentious and over-reliance on familiarity will breed stagnation.

    Invariance

    Principles of invariance have been a theme in scientific writing, and especially physics, since at least the early 20th century. The basic idea here is that good structures to look for are those independent of perspective, an idea that has featured earlier of course for example in Mill’s Methods of difference and agreement—methods that would be referred back to in the context of contrast and invariance. But as tends to be the case, there is a difference between something being a basic consideration and something being given weight. Principles of invariance have only been given weight in the wake of Einstein’s theories of relativity, which reduced everything to relations and were thereby fundamentally unchangeable, unable to be varied.As David Deutsch put it in 2009: “the search for hard-to-vary explanations is the origin of all progress”.

    An example here can be found in one of Einstein’s thought experiments. The one of a lab suspended in empty space is an example of a useful invariant observation. He imagined the absence of gravity and an experimenter free floating in the lab. — If now an entity pulls the lab upwards, accelerating uniformly, the experimenter would perceive the resulting force as gravity. The entity however would feel the work needed to accelerate the lab continuously. Through this experiment Einstein was able to equate gravitational and inertial mass; something unexplained by Newton’s laws, and an early but “powerful argument for a generalised postulate of relativity”.

    The feature, which suggests reality, is always some kind of invariance of a structure independent of the aspect, the projection.

    — Max Born, ‘Physical Reality’ (1953), 149 — as quoted by Weinert (2004)
    The discussion on invariance in physics is often had in the more specific context of symmetry. The Einstein example above, in the parlance of Mill would be an agreement between two values. In the context of invariance, it is a variable that remains unchanged through some kind of transformation or change in perspective. And discussion focused on symmetry would view the two perspectives as systems that share a relevant aspect and are therefore symmetrical.

    Related principles here are falsifiability and testability. The opposite of something being hard-to-vary are theories that resist falsification—a frustration that was expressed colourfully by Wolfgang Pauli as them being “not even wrong”. The importance of scientific theories to be falsifiable finds especial emphasis in the philosophy of Karl Popper. The broader view here is testability, since it includes the former and allows for additional practical considerations.

    Philosophy and discourse

    Philosophy of science looks at the underpinning logic of the scientific method, at what separates science from non-science, and the ethic that is implicit in science. There are basic assumptions, derived from philosophy by at least one prominent scientist, that form the base of the scientific method – namely, that reality is objective and consistent, that humans have the capacity to perceive reality accurately, and that rational explanations exist for elements of the real world. These assumptions from methodological naturalism form a basis on which science may be grounded. Logical positivist, empiricist, falsificationist, and other theories have criticized these assumptions and given alternative accounts of the logic of science, but each has also itself been criticized.

    There are several kinds of modern philosophical conceptualizations and attempts at definitions of the method of science.The one attempted by the unificationists, who argue for the existence of a unified definition that is useful (or at least ‘works’ in every context of science). The pluralists, arguing degrees of science being too fractured for a universal definition of its method to by useful. And those, who argue that the very attempt at definition is already detrimental to the free flow of ideas.

    Additionally, there have been views on the social framework in which science is done, and the impact of the sciences social environment on research. Also, there is ‘scientific method’ as popularised by Dewey in How We Think (1910) and Karl Pearson in Grammar of Science (1892), as used in fairly uncritical manner in education.

    Pluralism

    Scientific pluralism is a position within the philosophy of science that rejects various proposed unities of scientific method and subject matter. Scientific pluralists hold that science is not unified in one or more of the following ways: the metaphysics of its subject matter, the epistemology of scientific knowledge, or the research methods and models that should be used. Some pluralists believe that pluralism is necessary due to the nature of science. Others say that since scientific disciplines already vary in practice, there is no reason to believe this variation is wrong until a specific unification is empirically proven. Finally, some hold that pluralism should be allowed for normative reasons, even if unity were possible in theory.

    Unificationism

    Unificationism, in science, was a central tenet of logical positivism. Different logical positivists construed this doctrine in several different ways, e.g. as a reductionist thesis, that the objects investigated by the special sciences reduce to the objects of a common, putatively more basic domain of science, usually thought to be physics; as the thesis that all theories and results of the various sciences can or ought to be expressed in a common language or “universal slang”; or as the thesis that all the special sciences share a common scientific method.

    Development of the idea has been troubled by accelerated advancement in technology that has opened up many new ways to look at the world.

    The fact that the standards of scientific success shift with time does not only make the philosophy of science difficult; it also raises problems for the public understanding of science. We do not have a fixed scientific method to rally around and defend.

    — Steven Weinberg, 1995

    Epistemological anarchism

    Paul Feyerabend examined the history of science, and was led to deny that science is genuinely a methodological process. In his book Against Method he argued that no description of scientific method could possibly be broad enough to include all the approaches and methods used by scientists, and that there are no useful and exception-free methodological rules governing the progress of science. In essence, he said that for any specific method or norm of science, one can find a historic episode where violating it has contributed to the progress of science. He jokingly suggested that, if believers in the scientific method wish to express a single universally valid rule, it should be ‘anything goes’. As has been argued before him however, this is uneconomic; problem solvers, and researchers are to be prudent with their resources during their inquiry.

    A more general inference against formalised method has been found through research involving interviews with scientists regarding their conception of method. This research indicated that scientists frequently encounter difficulty in determining whether the available evidence supports their hypotheses. This reveals that there are no straightforward mappings between overarching methodological concepts and precise strategies to direct the conduct of research.

    Education
    In science education, the idea of a general and universal scientific method has been notably influential, and numerous studies (in the US) have shown that this framing of method often forms part of both students’ and teachers’ conception of science. This convention of traditional education has been argued against by scientists, as there is a consensus that educations’ sequential elements and unified view of scientific method do not reflect how scientists actually work. Major organizations of scientists such as the American Association for the Advancement of Science (AAAS) consider the sciences to be a part of the liberal arts traditions of learning and proper understating of science includes understanding of philosophy and history, not just science in isolation.

    How the sciences make knowledge has been taught in the context of “the” scientific method (singular) since the early 20th century. Various systems of education, including but not limited to the US, have taught the method of science as a process or procedure, structured as a definitive series of steps: observation, hypothesis, prediction, experiment.

    This version of the method of science has been a long-established standard in primary and secondary education, as well as the biomedical sciences. It has long been held to be an inaccurate idealisation of how some scientific inquiries are structured.

    The taught presentation of science had to defend demerits such as:

    it pays no regard to the social context of science,


    it suggests a singular methodology of deriving knowledge,


    it overemphasises experimentation,


    it oversimplifies science, giving the impression that following a scientific process automatically leads to knowledge,


    it gives the illusion of determination; that questions necessarily lead to some kind of answers and answers are preceded by (specific) questions,


    and, it holds that scientific theories arise from observed phenomena only.
    The scientific method no longer features in the standards for US education of 2013 (NGSS) that replaced those of 1996 (NRC). They, too, influenced international science education, and the standards measured for have shifted since from the singular hypothesis-testing method to a broader conception of scientific methods. These scientific methods, which are rooted in scientific practices and not epistemology, are described as the 3 dimensions of scientific and engineering practices, crosscutting concepts (interdisciplinary ideas), and disciplinary core ideas.

    The scientific method, as a result of simplified and universal explanations, is often held to have reached a kind of mythological status; as a tool for communication or, at best, an idealisation. Education’s approach was heavily influenced by John Dewey’s, How We Think (1910). Van der Ploeg (2016) indicated that Dewey’s views on education had long been used to further an idea of citizen education removed from “sound education”, claiming that references to Dewey in such arguments were undue interpretations (of Dewey).

    Sociology of knowledge

    The sociology of knowledge is a concept in the discussion around scientific method, claiming the underlying method of science to be sociological. King explains that sociology distinguishes here between the system of ideas that govern the sciences through an inner logic, and the social system in which those ideas arise.

    Thought collectives

    A perhaps accessible lead into what is claimed is Fleck’s thought, echoed in Kuhn’s concept of normal science. According to Fleck, scientists’ work is based on a thought-style, that cannot be rationally reconstructed. It gets instilled through the experience of learning, and science is then advanced based on a tradition of shared assumptions held by what he called thought collectives. Fleck also claims this phenomenon to be largely invisible to members of the group.

    Comparably, following the field research in an academic scientific laboratory by Latour and Woolgar, Karin Knorr Cetina has conducted a comparative study of two scientific fields (namely high energy physics and molecular biology) to conclude that the epistemic practices and reasonings within both scientific communities are different enough to introduce the concept of “epistemic cultures”, in contradiction with the idea that a so-called “scientific method” is unique and a unifying concept.

    Situated cognition and relativism

    On the idea of Fleck’s thought collectives sociologists built the concept of situated cognition: that the perspective of the researcher fundamentally affects their work; and, too, more radical views.

    Norwood Russell Hanson, alongside Thomas Kuhn and Paul Feyerabend, extensively explored the theory-laden nature of observation in science. Hanson introduced the concept in 1958, emphasizing that observation is influenced by the observer’s conceptual framework. He used the concept of gestalt to show how preconceptions can affect both observation and description, and illustrated this with examples like the initial rejection of Golgi bodies as an artefact of staining technique, and the differing interpretations of the same sunrise by Tycho Brahe and Johannes Kepler. Intersubjectivity led to different conclusions.

    Kuhn and Feyerabend acknowledged Hanson’s pioneering work, although Feyerabend’s views on methodological pluralism were more radical. Criticisms like those from Kuhn and Feyerabend prompted discussions leading to the development of the strong programme, a sociological approach that seeks to explain scientific knowledge without recourse to the truth or validity of scientific theories. It examines how scientific beliefs are shaped by social factors such as power, ideology, and interests.

    The postmodernist critiques of science have themselves been the subject of intense controversy. This ongoing debate, known as the science wars, is the result of conflicting values and assumptions between postmodernist and realist perspectives. Postmodernists argue that scientific knowledge is merely a discourse, devoid of any claim to fundamental truth. In contrast, realists within the scientific community maintain that science uncovers real and fundamental truths about reality. Many books have been written by scientists which take on this problem and challenge the assertions of the postmodernists while defending science as a legitimate way of deriving truth.

    Limits of method

    Role of chance in discovery

    Somewhere between 33% and 50% of all scientific discoveries are estimated to have been stumbled upon, rather than sought out. This may explain why scientists so often express that they were lucky. Scientists themselves in the 19th and 20th century acknowledged the role of fortunate luck or serendipity in discoveries. Louis Pasteur is credited with the famous saying that “Luck favours the prepared mind”, but some psychologists have begun to study what it means to be ‘prepared for luck’ in the scientific context. Research is showing that scientists are taught various heuristics that tend to harness chance and the unexpected. This is what Nassim Nicholas Taleb calls “Anti-fragility”; while some systems of investigation are fragile in the face of human error, human bias, and randomness, the scientific method is more than resistant or tough – it actually benefits from such randomness in many ways (it is anti-fragile). Taleb believes that the more anti-fragile the system, the more it will flourish in the real world.

    Psychologist Kevin Dunbar says the process of discovery often starts with researchers finding bugs in their experiments. These unexpected results lead researchers to try to fix what they think is an error in their method. Eventually, the researcher decides the error is too persistent and systematic to be a coincidence. The highly controlled, cautious, and curious aspects of the scientific method are thus what make it well suited for identifying such persistent systematic errors. At this point, the researcher will begin to think of theoretical explanations for the error, often seeking the help of colleagues across different domains of expertise.

    Relationship with statistics

    When the scientific method employs statistics as a key part of its arsenal, there are mathematical and practical issues that can have a deleterious effect on the reliability of the output of scientific methods. This is described in a popular 2005 scientific paper “Why Most Published Research Findings Are False” by John Ioannidis, which is considered foundational to the field of metascience. Much research in metascience seeks to identify poor use of statistics and improve its use, an example being the misuse of p-values.

    The particular points raised are statistical (“The smaller the studies conducted in a scientific field, the less likely the research findings are to be true” and “The greater the flexibility in designs, definitions, outcomes, and analytical modes in a scientific field, the less likely the research findings are to be true.”) and economical (“The greater the financial and other interests and prejudices in a scientific field, the less likely the research findings are to be true” and “The hotter a scientific field (with more scientific teams involved), the less likely the research findings are to be true.”) Hence: “Most research findings are false for most research designs and for most fields” and “As shown, the majority of modern biomedical research is operating in areas with very low pre- and poststudy probability for true findings.” However: “Nevertheless, most new discoveries will continue to stem from hypothesis-generating research with low or very low pre-study odds,” which means that new discoveries will come from research that, when that research started, had low or very low odds (a low or very low chance) of succeeding. Hence, if the scientific method is used to expand the frontiers of knowledge, research into areas that are outside the mainstream will yield the newest discoveries.

    Science of complex systems

    Science applied to complex systems can involve elements such as transdisciplinarity, systems theory, control theory, and scientific modelling.

    In general, the scientific method may be difficult to apply stringently to diverse, interconnected systems and large data sets. In particular, practices used within Big data, such as predictive analytics, may be considered to be at odds with the scientific method, as some of the data may have been stripped of the parameters which might be material in alternative hypotheses for an explanation; thus the stripped data would only serve to support the null hypothesis in the predictive analytics application. Fleck (1979), pp. 38–50 notes “a scientific discovery remains incomplete without considerations of the social practices that condition it”.

    Relationship with mathematics

    Science is the process of gathering, comparing, and evaluating proposed models against observables. A model can be a simulation, mathematical or chemical formula, or set of proposed steps. Science is like mathematics in that researchers in both disciplines try to distinguish what is known from what is unknown at each stage of discovery. Models, in both science and mathematics, need to be internally consistent and also ought to be falsifiable (capable of disproof). In mathematics, a statement need not yet be proved; at such a stage, that statement would be called a conjecture.

    Mathematical work and scientific work can inspire each other. For example, the technical concept of time arose in science, and timelessness was a hallmark of a mathematical topic. But today, the Poincaré conjecture has been proved using time as a mathematical concept in which objects can flow (see Ricci flow).

    Nevertheless, the connection between mathematics and reality (and so science to the extent it describes reality) remains obscure. Eugene Wigner’s paper, “The Unreasonable Effectiveness of Mathematics in the Natural Sciences”, is a very well-known account of the issue from a Nobel Prize-winning physicist. In fact, some observers (including some well-known mathematicians such as Gregory Chaitin, and others such as Lakoff and Núñez) have suggested that mathematics is the result of practitioner bias and human limitation (including cultural ones), somewhat like the post-modernist view of science.

    George Pólya’s work on problem solving, the construction of mathematical proofs, and heuristic show that the mathematical method and the scientific method differ in detail, while nevertheless resembling each other in using iterative or recursive steps.


    Mathematical method
    Scientific method
    1UnderstandingCharacterization from experience and observation
    2AnalysisHypothesis: a proposed explanation
    3SynthesisDeduction: prediction from the hypothesis
    4Review/ExtendTest and experiment

    In Pólya’s view, understanding involves restating unfamiliar definitions in your own words, resorting to geometrical figures, and questioning what we know and do not know already; analysis, which Pólya takes from Pappus, involves free and heuristic construction of plausible arguments, working backward from the goal, and devising a plan for constructing the proof; synthesis is the strict Euclidean exposition of step-by-step details of the proof; review involves reconsidering and re-examining the result and the path taken to it.

    Building on Pólya’s work, Imre Lakatos argued that mathematicians actually use contradiction, criticism, and revision as principles for improving their work.In like manner to science, where truth is sought, but certainty is not found, in Proofs and Refutations, what Lakatos tried to establish was that no theorem of informal mathematics is final or perfect. This means that, in non-axiomatic mathematics, we should not think that a theorem is ultimately true, only that no counterexample has yet been found. Once a counterexample, i.e. an entity contradicting/not explained by the theorem is found, we adjust the theorem, possibly extending the domain of its validity. This is a continuous way our knowledge accumulates, through the logic and process of proofs and refutations. (However, if axioms are given for a branch of mathematics, this creates a logical system —Wittgenstein 1921 Tractatus Logico-Philosophicus 5.13; Lakatos claimed that proofs from such a system were tautological, i.e. internally logically true, by rewriting forms, as shown by Poincaré, who demonstrated the technique of transforming tautologically true forms (viz. the Euler characteristic) into or out of forms from homology, or more abstractly, from homological algebra.

    Lakatos proposed an account of mathematical knowledge based on Polya’s idea of heuristics. In Proofs and Refutations, Lakatos gave several basic rules for finding proofs and counterexamples to conjectures. He thought that mathematical ‘thought experiments’ are a valid way to discover mathematical conjectures and proofs.

    Gauss, when asked how he came about his theorems, once replied “durch planmässiges Tattonieren” (through systematic palpable experimentation).

  • Science

    Science is a systematic discipline that builds and organises knowledge in the form of testable hypotheses and predictions about the universe.Modern science is typically divided into two or three major branches: the natural sciences (e.g., physics, chemistry, and biology), which study the physical world; and the social sciences (e.g., economics, psychology, and sociology), which study individuals and societies.Applied sciences are disciplines that use scientific knowledge for practical purposes, such as engineering and medicine. While sometimes referred to as the formal sciences, the study of logic, mathematics, and theoretical computer science (which study formal systems governed by axioms and rules) are typically regarded as separate because they rely on deductive reasoning instead of the scientific method or empirical evidence as their main methodology.

    The history of science spans the majority of the historical record, with the earliest identifiable predecessors to modern science dating to the Bronze Age in Egypt and Mesopotamia (c. 3000–1200 BCE). Their contributions to mathematics, astronomy, and medicine entered and shaped the Greek natural philosophy of classical antiquity, whereby formal attempts were made to provide explanations of events in the physical world based on natural causes, while further advancements, including the introduction of the Hindu–Arabic numeral system, were made during the Golden Age of India.Scientific research deteriorated in these regions after the fall of the Western Roman Empire during the Early Middle Ages (400–1000 CE), but in the Medieval renaissances (Carolingian Renaissance, Ottonian Renaissance and the Renaissance of the 12th century) scholarship flourished again. Some Greek manuscripts lost in Western Europe were preserved and expanded upon in the Middle East during the Islamic Golden Age, along with the later efforts of Byzantine Greek scholars who brought Greek manuscripts from the dying Byzantine Empire to Western Europe at the start of the Renaissance.

    The recovery and assimilation of Greek works and Islamic inquiries into Western Europe from the 10th to 13th centuries revived natural philosophy, which was later transformed by the Scientific Revolution that began in the 16th century as new ideas and discoveries departed from previous Greek conceptions and traditions. The scientific method soon played a greater role in knowledge creation and it was not until the 19th century that many of the institutional and professional features of science began to take shape, along with the changing of “natural philosophy” to “natural science”.

    New knowledge in science is advanced by research from scientists who are motivated by curiosity about the world and a desire to solve problems. Contemporary scientific research is highly collaborative and is usually done by teams in academic and research institutions, government agencies, and companies. The practical impact of their work has led to the emergence of science policies that seek to influence the scientific enterprise by prioritising the ethical and moral development of commercial products, armaments, health care, public infrastructure, and environmental protection.

    Etymology

    The word science has been used in Middle English since the 14th century in the sense of “the state of knowing”. The word was borrowed from the Anglo-Norman language as the suffix -cience, which was borrowed from the Latin word scientia, meaning “knowledge, awareness, understanding”, a noun derivative of sciens meaning “knowing”, itself the present active participle of sciō, “to know”.

    There are many hypotheses for science’s ultimate word origin. According to Michiel de Vaan, Dutch linguist and Indo-Europeanist, sciō may have its origin in the Proto-Italic language as *skije- or *skijo- meaning “to know”, which may originate from Proto-Indo-European language as *skh1-ie, *skh1-io, meaning “to incise”. The Lexikon der indogermanischen Verben proposed sciō is a back-formation of nescīre, meaning “to not know, be unfamiliar with”, which may derive from Proto-Indo-European *sekH- in Latin secāre, or *skh2-, from *sḱʰeh2(i)- meaning “to cut”.

    In the past, science was a synonym for “knowledge” or “study”, in keeping with its Latin origin. A person who conducted scientific research was called a “natural philosopher” or “man of science”. In 1834, William Whewell introduced the term scientist in a review of Mary Somerville’s book On the Connexion of the Physical Sciences, crediting it to “some ingenious gentleman” (possibly himself).

    History

    Early history

    Science has no single origin. Rather, scientific thinking emerged gradually over the course of tens of thousands of years, taking different forms around the world, and few details are known about the very earliest developments. Women likely played a central role in prehistoric science,as did religious rituals. Some scholars use the term “protoscience” to label activities in the past that resemble modern science in some but not all features;however, this label has also been criticised as denigrating, or too suggestive of presentism, thinking about those activities only in relation to modern categories.

    Direct evidence for scientific processes becomes clearer with the advent of writing systems in the Bronze Age civilisations of Ancient Egypt and Mesopotamia (bc.3000–1200 BCE), creating the earliest written records in the history of science.Although the words and concepts of “science” and “nature” were not part of the conceptual landscape at the time, the ancient Egyptians and Mesopotamians made contributions that would later find a place in Greek and medieval science: mathematics, astronomy, and medicine.From the 3rd millennium BCE, the ancient Egyptians developed a non-positional decimal numbering system,[ solved practical problems using geometry, and developed a calendar. Their healing therapies involved drug treatments and the supernatural, such as prayers, incantations, and rituals.

    The ancient Mesopotamians used knowledge about the properties of various natural chemicals for manufacturing pottery, faience, glass, soap, metals, lime plaster, and waterproofing. They studied animal physiology, anatomy, behaviour, and astrology for divinatory purposes. The Mesopotamians had an intense interest in medicine and the earliest medical prescriptions appeared in Sumerian during the Third Dynasty of Ur.They seem to have studied scientific subjects which had practical or religious applications and had little interest in satisfying curiosity.

    Classical antiquity

    In classical antiquity, there is no real ancient analogue of a modern scientist. Instead, well-educated, usually upper-class, and almost universally male individuals performed various investigations into nature whenever they could afford the time. Before the invention or discovery of the concept of phusis or nature by the pre-Socratic philosophers, the same words tend to be used to describe the natural “way” in which a plant grows, and the “way” in which, for example, one tribe worships a particular god. For this reason, it is claimed that these men were the first philosophers in the strict sense and the first to clearly distinguish “nature” and “convention”.

    The early Greek philosophers of the Milesian school, which was founded by Thales of Miletus and later continued by his successors Anaximander and Anaximenes, were the first to attempt to explain natural phenomena without relying on the supernatural. The Pythagoreans developed a complex number philosophy467–468 and contributed significantly to the development of mathematical science.The theory of atoms was developed by the Greek philosopher Leucippus and his student Democritus. Later, Epicurus would develop a full natural cosmology based on atomism, and would adopt a “canon” (ruler, standard) which established physical criteria or standards of scientific truth. The Greek doctor Hippocrates established the tradition of systematic medical science and is known as “The Father of Medicine”.

    A turning point in the history of early philosophical science was Socrates’ example of applying philosophy to the study of human matters, including human nature, the nature of political communities, and human knowledge itself. The Socratic method as documented by Plato’s dialogues is a dialectic method of hypothesis elimination: better hypotheses are found by steadily identifying and eliminating those that lead to contradictions. The Socratic method searches for general commonly-held truths that shape beliefs and scrutinises them for consistency. Socrates criticised the older type of study of physics as too purely speculative and lacking in self-criticism.

    In the 4th century BCE, Aristotle created a systematic programme of teleological philosophy.In the 3rd century BCE, Greek astronomer Aristarchus of Samos was the first to propose a heliocentric model of the universe, with the Sun at the centre and all the planets orbiting it. Aristarchus’s model was widely rejected because it was believed to violate the laws of physics, while Ptolemy’s Almagest, which contains a geocentric description of the Solar System, was accepted through the early Renaissance instead.The inventor and mathematician Archimedes of Syracuse made major contributions to the beginnings of calculus. Pliny the Elder was a Roman writer and polymath, who wrote the seminal encyclopaedia Natural History.

    Positional notation for representing numbers likely emerged between the 3rd and 5th centuries CE along Indian trade routes. This numeral system made efficient arithmetic operations more accessible and would eventually become standard for mathematics worldwide.

    Middle Ages

    Due to the collapse of the Western Roman Empire, the 5th century saw an intellectual decline, with knowledge of classical Greek conceptions of the world deteriorating in Western Europe. 194 Latin encyclopaedists of the period such as Isidore of Seville preserved the majority of general ancient knowledge. In contrast, because the Byzantine Empire resisted attacks from invaders, they were able to preserve and improve prior learning.159 John Philoponus, a Byzantine scholar in the 6th century, started to question Aristotle’s teaching of physics, introducing the theory of impetus.307,311,363,402 His criticism served as an inspiration to medieval scholars and Galileo Galilei, who extensively cited his works ten centuries later.307–308

    During late antiquity and the Early Middle Ages, natural phenomena were mainly examined via the Aristotelian approach. The approach includes Aristotle’s four causes: material, formal, moving, and final cause. Many Greek classical texts were preserved by the Byzantine Empire and Arabic translations were done by groups such as the Nestorians and the Monophysites. Under the Abbasids, these Arabic translations were later improved and developed by Arabic scientists. By the 6th and 7th centuries, the neighbouring Sasanian Empire established the medical Academy of Gondishapur, which was considered by Greek, Syriac, and Persian physicians as the most important medical hub of the ancient world.

    Islamic study of Aristotelianism flourished in the House of Wisdom established in the Abbasid capital of Baghdad, Iraq and the flourished until the Mongol invasions in the 13th century. Ibn al-Haytham, better known as Alhazen, used controlled experiments in his optical study. Avicenna’s compilation of The Canon of Medicine, a medical encyclopaedia, is considered to be one of the most important publications in medicine and was used until the 18th century.

    By the 11th century most of Europe had become Christian,204 and in 1088, the University of Bologna emerged as the first university in Europe. As such, demand for Latin translation of ancient and scientific texts grew,204 a major contributor to the Renaissance of the 12th century. Renaissance scholasticism in western Europe flourished, with experiments done by observing, describing, and classifying subjects in nature. In the 13th century, medical teachers and students at Bologna began opening human bodies, leading to the first anatomy textbook based on human dissection by Mondino de Luzzi.

    Renaissance

    New developments in optics played a role in the inception of the Renaissance, both by challenging long-held metaphysical ideas on perception, as well as by contributing to the improvement and development of technology such as the camera obscura and the telescope. At the start of the Renaissance, Roger Bacon, Vitello, and John Peckham each built up a scholastic ontology upon a causal chain beginning with sensation, perception, and finally apperception of the individual and universal forms of Aristotle. Book A model of vision later known as perspectivism was exploited and studied by the artists of the Renaissance. This theory uses only three of Aristotle’s four causes: formal, material, and final.

    In the 16th century, Nicolaus Copernicus formulated a heliocentric model of the Solar System, stating that the planets revolve around the Sun, instead of the geocentric model where the planets and the Sun revolve around the Earth. This was based on a theorem that the orbital periods of the planets are longer as their orbs are farther from the centre of motion, which he found not to agree with Ptolemy’s model.

    Johannes Kepler and others challenged the notion that the only function of the eye is perception, and shifted the main focus in optics from the eye to the propagation of light. Kepler is best known, however, for improving Copernicus’ heliocentric model through the discovery of Kepler’s laws of planetary motion. Kepler did not reject Aristotelian metaphysics and described his work as a search for the Harmony of the Spheres. Galileo had made significant contributions to astronomy, physics and engineering. However, he became persecuted after Pope Urban VIII sentenced him for writing about the heliocentric model.

    The printing press was widely used to publish scholarly arguments, including some that disagreed widely with contemporary ideas of nature. Francis Bacon and René Descartes published philosophical arguments in favour of a new type of non-Aristotelian science. Bacon emphasised the importance of experiment over contemplation, questioned the Aristotelian concepts of formal and final cause, promoted the idea that science should study the laws of nature and the improvement of all human life. Descartes emphasised individual thought and argued that mathematics rather than geometry should be used to study nature.

    Age of Enlightenment

    At the start of the Age of Enlightenment, Isaac Newton formed the foundation of classical mechanics by his Philosophiæ Naturalis Principia Mathematica, greatly influencing future physicists. Gottfried Wilhelm Leibniz incorporated terms from Aristotelian physics, now used in a new non-teleological way. This implied a shift in the view of objects: objects were now considered as having no innate goals. Leibniz assumed that different types of things all work according to the same general laws of nature, with no special formal or final causes.

    During this time the declared purpose and value of science became producing wealth and inventions that would improve human lives, in the materialistic sense of having more food, clothing, and other things. In Bacon’s words, “the real and legitimate goal of sciences is the endowment of human life with new inventions and riches”, and he discouraged scientists from pursuing intangible philosophical or spiritual ideas, which he believed contributed little to human happiness beyond “the fume of subtle, sublime or pleasing”.

    Science during the Enlightenment was dominated by scientific societies and academies,which had largely replaced universities as centres of scientific research and development. Societies and academies were the backbones of the maturation of the scientific profession. Another important development was the popularisation of science among an increasingly literate population. Enlightenment philosophers turned to a few of their scientific predecessors – Galileo, Kepler, Boyle, and Newton principally – as the guides to every physical and social field of the day.

    The 18th century saw significant advancements in the practice of medicine and physics; the development of biological taxonomy by Carl Linnaeus; a new understanding of magnetism and electricity; and the maturation of chemistry as a discipline. Ideas on human nature, society, and economics evolved during the Enlightenment. Hume and other Scottish Enlightenment thinkers developed A Treatise of Human Nature, which was expressed historically in works by authors including James Burnett, Adam Ferguson, John Millar and William Robertson, all of whom merged a scientific study of how humans behaved in ancient and primitive cultures with a strong awareness of the determining forces of modernity. Modern sociology largely originated from this movement. In 1776, Adam Smith published The Wealth of Nations, which is often considered the first work on modern economics.

    19th century

    During the 19th century, many distinguishing characteristics of contemporary modern science began to take shape. These included the transformation of the life and physical sciences; the frequent use of precision instruments; the emergence of terms such as “biologist”, “physicist”, and “scientist”; an increased professionalisation of those studying nature; scientists gaining cultural authority over many dimensions of society; the industrialisation of numerous countries; the thriving of popular science writings; and the emergence of science journals. During the late 19th century, psychology emerged as a separate discipline from philosophy when Wilhelm Wundt founded the first laboratory for psychological research in 1879.

    During the mid-19th century Charles Darwin and Alfred Russel Wallace independently proposed the theory of evolution by natural selection in 1858, which explained how different plants and animals originated and evolved. Their theory was set out in detail in Darwin’s book On the Origin of Species, published in 1859. Separately, Gregor Mendel presented his paper, “Experiments on Plant Hybridisation” in 1865, which outlined the principles of biological inheritance, serving as the basis for modern genetics.

    Early in the 19th century John Dalton suggested the modern atomic theory, based on Democritus’s original idea of indivisible particles called atoms. The laws of conservation of energy, conservation of momentum and conservation of mass suggested a highly stable universe where there could be little loss of resources. However, with the advent of the steam engine and the Industrial Revolution there was an increased understanding that not all forms of energy have the same energy qualities, the ease of conversion to useful work or to another form of energy. This realisation led to the development of the laws of thermodynamics, in which the free energy of the universe is seen as constantly declining: the entropy of a closed universe increases over time.

    The electromagnetic theory was established in the 19th century by the works of Hans Christian Ørsted, André-Marie Ampère, Michael Faraday, James Clerk Maxwell, Oliver Heaviside, and Heinrich Hertz. The new theory raised questions that could not easily be answered using Newton’s framework. The discovery of X-rays inspired the discovery of radioactivity by Henri Becquerel and Marie Curie in 1896, Marie Curie then became the first person to win two Nobel Prizes. In the next year came the discovery of the first subatomic particle, the electron.

    20th century

    In the first half of the century the development of antibiotics and artificial fertilisers improved human living standards globally. Harmful environmental issues such as ozone depletion, ocean acidification, eutrophication, and climate change came to the public’s attention and caused the onset of environmental studies.

    During this period scientific experimentation became increasingly larger in scale and funding. The extensive technological innovation stimulated by World War I, World War II, and the Cold War led to competitions between global powers, such as the Space Race and nuclear arms race. Substantial international collaborations were also made, despite armed conflicts.

    In the late 20th century active recruitment of women and elimination of sex discrimination greatly increased the number of women scientists, but large gender disparities remained in some fields. The discovery of the cosmic microwave background in 1964 led to a rejection of the steady-state model of the universe in favour of the Big Bang theory of Georges Lemaître.

    The century saw fundamental changes within science disciplines. Evolution became a unified theory in the early 20th-century when the modern synthesis reconciled Darwinian evolution with classical genetics. Albert Einstein’s theory of relativity and the development of quantum mechanics complement classical mechanics to describe physics in extreme length, time and gravity. Widespread use of integrated circuits in the last quarter of the 20th century combined with communications satellites led to a revolution in information technology and the rise of the global internet and mobile computing, including smartphones. The need for mass systematisation of long, intertwined causal chains and large amounts of data led to the rise of the fields of systems theory and computer-assisted scientific modelling.

    21st century

    The Human Genome Project was completed in 2003 by identifying and mapping all of the genes of the human genome. The first induced pluripotent human stem cells were made in 2006, allowing adult cells to be transformed into stem cells and turn into any cell type found in the body. With the affirmation of the Higgs boson discovery in 2013, the last particle predicted by the Standard Model of particle physics was found. In 2015, gravitational waves, predicted by general relativity a century before, were first observed. In 2019, the international collaboration Event Horizon Telescope presented the first direct image of a black hole’s accretion disc.

    Branches

    Modern science is commonly divided into three major branches: natural science, social science, and formal science. Each of these branches comprises various specialised yet overlapping scientific disciplines that often possess their own nomenclature and expertise. Both natural and social sciences are empirical sciences, as their knowledge is based on empirical observations and is capable of being tested for its validity by other researchers working under the same conditions.

    Natural science

    Natural science is the study of the physical world. It can be divided into two main branches: life science and physical science. These two branches may be further divided into more specialised disciplines. For example, physical science can be subdivided into physics, chemistry, astronomy, and earth science. Modern natural science is the successor to the natural philosophy that began in Ancient Greece. Galileo, Descartes, Bacon, and Newton debated the benefits of using approaches that were more mathematical and more experimental in a methodical way. Still, philosophical perspectives, conjectures, and presuppositions, often overlooked, remain necessary in natural science. Systematic data collection, including discovery science, succeeded natural history, which emerged in the 16th century by describing and classifying plants, animals, minerals, and other biotic beings.Today, “natural history” suggests observational descriptions aimed at popular audiences.

    Social science

    Social science is the study of human behaviour and the functioning of societies. It has many disciplines that include, but are not limited to anthropology, economics, history, human geography, political science, psychology, and sociology. In the social sciences, there are many competing theoretical perspectives, many of which are extended through competing research programmes such as the functionalists, conflict theorists, and interactionists in sociology. Due to the limitations of conducting controlled experiments involving large groups of individuals or complex situations, social scientists may adopt other research methods such as the historical method, case studies, and cross-cultural studies. Moreover, if quantitative information is available, social scientists may rely on statistical approaches to better understand social relationships and processes.

    Formal science

    Formal science is an area of study that generates knowledge using formal systems. A formal system is an abstract structure used for inferring theorems from axioms according to a set of rules. It includes mathematics,systems theory, and theoretical computer science. The formal sciences share similarities with the other two branches by relying on objective, careful, and systematic study of an area of knowledge. They are, however, different from the empirical sciences as they rely exclusively on deductive reasoning, without the need for empirical evidence, to verify their abstract concepts. The formal sciences are therefore a priori disciplines and because of this, there is disagreement on whether they constitute a science. Nevertheless, the formal sciences play an important role in the empirical sciences. Calculus, for example, was initially invented to understand motion in physics. Natural and social sciences that rely heavily on mathematical applications include mathematical physics, chemistry, biology, finance, and economics.

    Applied science

    Applied science is the use of the scientific method and knowledge to attain practical goals and includes a broad range of disciplines such as engineering and medicine. Engineering is the use of scientific principles to invent, design and build machines, structures and technologies. Science may contribute to the development of new technologies. Medicine is the practice of caring for patients by maintaining and restoring health through the prevention, diagnosis, and treatment of injury or disease.

    Basic sciences

    The applied sciences are often contrasted with the basic sciences, which are focused on advancing scientific theories and laws that explain and predict events in the natural world.

    Blue skies science

    Blue skies research, also called blue sky science, is scientific research in domains where “real-world” applications are not immediately apparent. It has been defined as “research without a clear goal” and “curiosity-driven science”. Proponents of this mode of science argue that unanticipated scientific breakthroughs are sometimes more valuable than the outcomes of agenda-driven research, heralding advances in genetics and stem cell biology as examples of unforeseen benefits of research that was originally seen as purely theoretical in scope. Because of the inherently uncertain return on investment, blue-sky projects are sometimes politically and commercially unpopular and tend to lose funding to research perceived as being more reliably profitable or practical.

    Computational science

    Computational science applies computer simulations to science, enabling a better understanding of scientific problems than formal mathematics alone can achieve. The use of machine learning and artificial intelligence is becoming a central feature of computational contributions to science, for example in agent-based computational economics, random forests, topic modeling and various forms of prediction. However, machines alone rarely advance knowledge as they require human guidance and capacity to reason; and they can introduce bias against certain social groups or sometimes underperform against humans.

    Interdisciplinary science

    Interdisciplinary science involves the combination of two or more disciplines into one, such as bioinformatics, a combination of biology and computer science or cognitive sciences. The concept has existed since the ancient Greek period and it became popular again in the 20th century.

    Scientific research

    Scientific research can be labelled as either basic or applied research. Basic research is the search for knowledge and applied research is the search for solutions to practical problems using this knowledge. Most understanding comes from basic research, though sometimes applied research targets specific practical problems. This leads to technological advances that were not previously imaginable.

    Scientific method

    Scientific research involves using the scientific method, which seeks to objectively explain the events of nature in a reproducible way. Scientists usually take for granted a set of basic assumptions that are needed to justify the scientific method: there is an objective reality shared by all rational observers; this objective reality is governed by natural laws; these laws were discovered by means of systematic observation and experimentation. Mathematics is essential in the formation of hypotheses, theories, and laws, because it is used extensively in quantitative modelling, observing, and collecting measurements. Statistics is used to summarise and analyse data, which allows scientists to assess the reliability of experimental results.

    In the scientific method an explanatory thought experiment or hypothesis is put forward as an explanation using parsimony principles and is expected to seek consilience – fitting with other accepted facts related to an observation or scientific question. This tentative explanation is used to make falsifiable predictions, which are typically posted before being tested by experimentation. Disproof of a prediction is evidence of progress.4–5 Experimentation is especially important in science to help establish causal relationships to avoid the correlation fallacy, though in some sciences such as astronomy or geology, a predicted observation might be more appropriate.

    When a hypothesis proves unsatisfactory it is modified or discarded. If the hypothesis survives testing, it may become adopted into the framework of a scientific theory, a validly reasoned, self-consistent model or framework for describing the behaviour of certain natural events. A theory typically describes the behaviour of much broader sets of observations than a hypothesis; commonly, a large number of hypotheses can be logically bound together by a single theory. Thus, a theory is a hypothesis explaining various other hypotheses. In that vein, theories are formulated according to most of the same scientific principles as hypotheses. Scientists may generate a model, an attempt to describe or depict an observation in terms of a logical, physical or mathematical representation, and to generate new hypotheses that can be tested by experimentation.

    While performing experiments to test hypotheses, scientists may have a preference for one outcome over another. Eliminating the bias can be achieved through transparency, careful experimental design, and a thorough peer review process of the experimental results and conclusions.After the results of an experiment are announced or published, it is normal practice for independent researchers to double-check how the research was performed, and to follow up by performing similar experiments to determine how dependable the results might be. Taken in its entirety, the scientific method allows for highly creative problem solving while minimising the effects of subjective and confirmation bias.Intersubjective verifiability, the ability to reach a consensus and reproduce results, is fundamental to the creation of all scientific knowledge.

    Scientific literature

    Scientific research is published in a range of literature. Scientific journals communicate and document the results of research carried out in universities and various other research institutions, serving as an archival record of science. The first scientific journals, Journal des sçavans followed by Philosophical Transactions, began publication in 1665. Since that time the total number of active periodicals has steadily increased. In 1981, one estimate for the number of scientific and technical journals in publication was 11,500.

    Most scientific journals cover a single scientific field and publish the research within that field; the research is normally expressed in the form of a scientific paper. Science has become so pervasive in modern societies that it is considered necessary to communicate the achievements, news, and ambitions of scientists to a wider population.

    Challenges

    The replication crisis is an ongoing methodological crisis that affects parts of the social and life sciences. In subsequent investigations, the results of many scientific studies have been proven to be unrepeatable. The crisis has long-standing roots; the phrase was coined in the early 2010s as part of a growing awareness of the problem. The replication crisis represents an important body of research in metascience, which aims to improve the quality of all scientific research while reducing waste.

    An area of study or speculation that masquerades as science in an attempt to claim legitimacy that it would not otherwise be able to achieve is sometimes referred to as pseudoscience, fringe science, or junk science. Physicist Richard Feynman coined the term “cargo cult science” for cases in which researchers believe, and at a glance, look like they are doing science but lack the honesty to allow their results to be rigorously evaluated. Various types of commercial advertising, ranging from hype to fraud, may fall into these categories. Science has been described as “the most important tool” for separating valid claims from invalid ones.

    There can also be an element of political bias or ideological bias on all sides of scientific debates. Sometimes, research may be characterised as “bad science”, research that may be well-intended but is incorrect, obsolete, incomplete, or over-simplified expositions of scientific ideas. The term scientific misconduct refers to situations such as where researchers have intentionally misrepresented their published data or have purposely given credit for a discovery to the wrong person.

    Philosophy of science

    There are different schools of thought in the philosophy of science. The most popular position is empiricism, which holds that knowledge is created by a process involving observation; scientific theories generalise observations. Empiricism generally encompasses inductivism, a position that explains how general theories can be made from the finite amount of empirical evidence available. Many versions of empiricism exist, with the predominant ones being Bayesianism and the hypothetico-deductive method.

    Empiricism has stood in contrast to rationalism, the position originally associated with Descartes, which holds that knowledge is created by the human intellect, not by observation. Critical rationalism is a contrasting 20th-century approach to science, first defined by Austrian-British philosopher Karl Popper. Popper rejected the way that empiricism describes the connection between theory and observation. He claimed that theories are not generated by observation, but that observation is made in the light of theories, and that the only way theory A can be affected by observation is after theory A were to conflict with observation, but theory B were to survive the observation. Popper proposed replacing verifiability with falsifiability as the landmark of scientific theories, replacing induction with falsification as the empirical method. Popper further claimed that there is actually only one universal method, not specific to science: the negative method of criticism, trial and error, covering all products of the human mind, including science, mathematics, philosophy, and art.

    Another approach, instrumentalism, emphasises the utility of theories as instruments for explaining and predicting phenomena. It views scientific theories as black boxes, with only their input (initial conditions) and output (predictions) being relevant. Consequences, theoretical entities, and logical structure are claimed to be things that should be ignored. Close to instrumentalism is constructive empiricism, according to which the main criterion for the success of a scientific theory is whether what it says about observable entities is true.

    Thomas Kuhn argued that the process of observation and evaluation takes place within a paradigm, a logically consistent “portrait” of the world that is consistent with observations made from its framing. He characterised normal science as the process of observation and “puzzle solving”, which takes place within a paradigm, whereas revolutionary science occurs when one paradigm overtakes another in a paradigm shift. Each paradigm has its own distinct questions, aims, and interpretations. The choice between paradigms involves setting two or more “portraits” against the world and deciding which likeness is most promising. A paradigm shift occurs when a significant number of observational anomalies arise in the old paradigm and a new paradigm makes sense of them. That is, the choice of a new paradigm is based on observations, even though those observations are made against the background of the old paradigm. For Kuhn, acceptance or rejection of a paradigm is a social process as much as a logical process. Kuhn’s position, however, is not one of relativism.

    Another approach often cited in debates of scientific scepticism against controversial movements like “creation science” is methodological naturalism. Naturalists maintain that a difference should be made between natural and supernatural, and science should be restricted to natural explanations. Methodological naturalism maintains that science requires strict adherence to empirical study and independent verification.

    Scientific community

    The scientific community is a network of interacting scientists who conduct scientific research. The community consists of smaller groups working in scientific fields. By having peer review, through discussion and debate within journals and conferences, scientists maintain the quality of research methodology and objectivity when interpreting results.

    Scientists

    Scientists are individuals who conduct scientific research to advance knowledge in an area of interest. Scientists may exhibit a strong curiosity about reality and a desire to apply scientific knowledge for the benefit of public health, nations, the environment, or industries; other motivations include recognition by peers and prestige. In modern times, many scientists study within specific areas of science in academic institutions, often obtaining advanced degrees in the process. Many scientists pursue careers in various fields such as academia, industry, government, and nonprofit organisations.

    Science has historically been a male-dominated field, with notable exceptions. Women have faced considerable discrimination in science, much as they have in other areas of male-dominated societies. For example, women were frequently passed over for job opportunities and denied credit for their work. The achievements of women in science have been attributed to the defiance of their traditional role as labourers within the domestic sphere.

    Learned societies

    Learned societies for the communication and promotion of scientific thought and experimentation have existed since the Renaissance. Many scientists belong to a learned society that promotes their respective scientific discipline, profession, or group of related disciplines. Membership may either be open to all, require possession of scientific credentials, or conferred by election.Most scientific societies are nonprofit organisations, and many are professional associations. Their activities typically include holding regular conferences for the presentation and discussion of new research results and publishing or sponsoring academic journals in their discipline. Some societies act as professional bodies, regulating the activities of their members in the public interest, or the collective interest of the membership.

    The professionalisation of science, begun in the 19th century, was partly enabled by the creation of national distinguished academies of sciences such as the Italian Accademia dei Lincei in 1603, the British Royal Society in 1660, the French Academy of Sciences in 1666, the American National Academy of Sciences in 1863, the German Kaiser Wilhelm Society in 1911, and the Chinese Academy of Sciences in 1949. International scientific organisations, such as the International Science Council, are devoted to international cooperation for science advancement.

    Awards

    Science awards are usually given to individuals or organisations that have made significant contributions to a discipline. They are often given by prestigious institutions; thus, it is considered a great honour for a scientist receiving them. Since the early Renaissance, scientists have often been awarded medals, money, and titles. The Nobel Prize, a widely regarded prestigious award, is awarded annually to those who have achieved scientific advances in the fields of medicine, physics, and chemistry.

    Society

    Funding and policies

    Funding of science is often through a competitive process in which potential research projects are evaluated and only the most promising receive funding. Such processes, which are run by government, corporations, or foundations, allocate scarce funds. Total research funding in most developed countries is between 1.5% and 3% of GDP. In the OECD, around two-thirds of research and development in scientific and technical fields is carried out by industry, and 20% and 10%, respectively, by universities and government. The government funding proportion in certain fields is higher, and it dominates research in social science and the humanities. In less developed nations, the government provides the bulk of the funds for their basic scientific research.

    Many governments have dedicated agencies to support scientific research, such as the National Science Foundation in the United States, the National Scientific and Technical Research Council in Argentina, Commonwealth Scientific and Industrial Research Organisation in Australia, National Centre for Scientific Research in France, the Max Planck Society in Germany, and National Research Council in Spain. In commercial research and development, all but the most research-orientated corporations focus more heavily on near-term commercialisation possibilities than research driven by curiosity.

    Science policy is concerned with policies that affect the conduct of the scientific enterprise, including research funding, often in pursuance of other national policy goals such as technological innovation to promote commercial product development, weapons development, health care, and environmental monitoring. Science policy sometimes refers to the act of applying scientific knowledge and consensus to the development of public policies. In accordance with public policy being concerned about the well-being of its citizens, science policy’s goal is to consider how science and technology can best serve the public. Public policy can directly affect the funding of capital equipment and intellectual infrastructure for industrial research by providing tax incentives to those organisations that fund research.

    Education and awareness

    Science education for the general public is embedded in the school curriculum, and is supplemented by online pedagogical content (for example, YouTube and Khan Academy), museums, and science magazines and blogs. Major organisations of scientists such as the American Association for the Advancement of Science (AAAS) consider the sciences to be a part of the liberal arts traditions of learning, along with philosophy and history. Scientific literacy is chiefly concerned with an understanding of the scientific method, units and methods of measurement, empiricism, a basic understanding of statistics (correlations, qualitative versus quantitative observations, aggregate statistics), and a basic understanding of core scientific fields such as physics, chemistry, biology, ecology, geology, and computation. As a student advances into higher stages of formal education, the curriculum becomes more in depth. Traditional subjects usually included in the curriculum are natural and formal sciences, although recent movements include social and applied science as well.

    The mass media face pressures that can prevent them from accurately depicting competing scientific claims in terms of their credibility within the scientific community as a whole. Determining how much weight to give different sides in a scientific debate may require considerable expertise regarding the matter. Few journalists have real scientific knowledge, and even beat reporters who are knowledgeable about certain scientific issues may be ignorant about other scientific issues that they are suddenly asked to cover.

    Science magazines such as New Scientist, Science & Vie, and Scientific American cater to the needs of a much wider readership and provide a non-technical summary of popular areas of research, including notable discoveries and advances in certain fields of research. The science fiction genre, primarily speculative fiction, can transmit the ideas and methods of science to the general public. Recent efforts to intensify or develop links between science and non-scientific disciplines, such as literature or poetry, include the Creative Writing Science resource developed through the Royal Literary Fund.

    Anti-science attitudes

    While the scientific method is broadly accepted in the scientific community, some fractions of society reject certain scientific positions or are sceptical about science. Examples are the common notion that COVID-19 is not a major health threat to the US (held by 39% of Americans in August 2021) or the belief that climate change is not a major threat to the US (also held by 40% of Americans, in late 2019 and early 2020). Psychologists have pointed to four factors driving rejection of scientific results:

    Scientific authorities are sometimes seen as inexpert, untrustworthy, or biased.
    Some marginalised social groups hold anti-science attitudes, in part because these groups have often been exploited in unethical experiments.
    Messages from scientists may contradict deeply held existing beliefs or morals.
    The delivery of a scientific message may not be appropriately targeted to a recipient’s learning style.
    Anti-science attitudes often seem to be caused by fear of rejection in social groups. For instance, climate change is perceived as a threat by only 22% of Americans on the right side of the political spectrum, but by 85% on the left.That is, if someone on the left would not consider climate change as a threat, this person may face contempt and be rejected in that social group. In fact, people may rather deny a scientifically accepted fact than lose or jeopardise their social status.

    Politics

    Attitudes towards science are often determined by political opinions and goals. Government, business and advocacy groups have been known to use legal and economic pressure to influence scientific researchers. Many factors can act as facets of the politicisation of science such as anti-intellectualism, perceived threats to religious beliefs, and fear for business interests. Politicisation of science is usually accomplished when scientific information is presented in a way that emphasises the uncertainty associated with the scientific evidence. Tactics such as shifting conversation, failing to acknowledge facts, and capitalising on doubt of scientific consensus have been used to gain more attention for views that have been undermined by scientific evidence. Examples of issues that have involved the politicisation of science include the global warming controversy, health effects of pesticides, and health effects of tobacco.